Layered double hydroxides (LDHs), also known as anionic clays or hydrotalcite-like compounds, are two-dimensional, synthetic or naturally occurring layered materials composed of positively charged brucite-like (Mg(OH)2) layers formed by divalent (M2+, e.g., Mg2+, Zn2+, Ca2+) and trivalent (M3+, e.g., Al3+, Cr3+) metal cations in hydroxide octahedral coordination, with charge-balancing interlayer anions (e.g., CO32-, NO3-, Cl-) and water molecules providing structural stability and hydration.[1][2][3]The general chemical formula for LDHs is [M2+1-xM3+x(OH)2]x+(An-)x/n·yH2O, where x represents the molar ratio of M3+ to the total metal content (typically 0.2 ≤ x ≤ 0.33, though ranging from 0.1 to 0.5), allowing tunable composition and properties through variation in metal cations and interlayer anions.[1][2][3] First identified as the mineral hydrotalcite in 1842 by Hochstetter in Norway, LDHs were synthesized in laboratories starting in the 1940s by Feitknecht and gained commercial interest after a 1970 German patent for hydrotalcite-based antacids, with extensive research accelerating in the 1990s for advanced applications.[2][4][5]LDHs exhibit distinctive properties such as high specific surface area (often >100 m²/g), strong anion exchange capacity, thermal stability up to 400–500°C, chemical stability across a wide pH range, biocompatibility, and the unique "memory effect" where calcined LDHs reconstruct their layered structure upon rehydration or anion exposure.[1][2][3] These attributes stem from their tunable interlayer spacing (0.78–0.80 nm basal spacing) and high charge density, enabling pH-dependent solubility and self-healing behaviors in composites.[2]Common synthesis methods include co-precipitation under controlled pH (9–12) and temperature (50–90°C) for high yields, hydrothermal treatment up to 180°C for morphology control, and sol-gel processes using metal alkoxides for high-purity materials with large surface areas; greener approaches like mechanochemical grinding and ureahydrolysis have emerged to minimize waste.[1][2][3]Due to their versatility, LDHs find applications in catalysis (e.g., biodiesel transesterification and water splitting), environmental remediation (e.g., adsorption of dyes like methyl orange at 185 mg/g and heavy metals), drug delivery systems (e.g., controlled release of curcumin or in antacid formulations like TALCID®), polymer nanocomposites for flame retardancy and gas barriers, energy storage, photocatalysis, and even food packaging and piezoelectric devices. Recent research as of 2025 has expanded LDH applications to piezocatalysis for tumor therapy and regenerative nanomedicine.[1][2][3][6][7]
Composition and Structure
General Formula and Composition
Layered double hydroxides (LDHs) are a class of anionic clays characterized by their tunable chemical composition, derived from the partial substitution of divalent cations in brucite-like hydroxide layers with trivalent cations, resulting in positively charged sheets balanced by interlayer anions and water molecules.[8] The general formula for most LDHs is [ \mathrm{M}^{2+}_{1-x} \mathrm{M}^{3+}_x (\mathrm{OH})_2 ]^{x+} [ \mathrm{A}^{n-}_{x/n} \cdot m \mathrm{H}_2\mathrm{O} ]^{x-}, where \mathrm{M}^{2+} represents divalent metal cations such as \mathrm{Mg}^{2+}, \mathrm{Zn}^{2+}, or \mathrm{Ca}^{2+}; \mathrm{M}^{3+} denotes trivalent cations like \mathrm{Al}^{3+}, \mathrm{Fe}^{3+}, or \mathrm{Cr}^{3+}; \mathrm{A}^{n-} are interlayer anions including \mathrm{CO}_3^{2-}, \mathrm{Cl}^{-}, \mathrm{NO}_3^{-}, or \mathrm{SO}_4^{2-}; x is the fraction of trivalent cations, typically ranging from 0.2 to 0.33 to ensure structural stability; and m indicates the variable water content in the interlayer region.[6] This formula reflects the positive charge density on the hydroxide layers, which arises from the isomorphous substitution and is precisely balanced by the charge of the anions.[9]The value of x plays a critical role in determining the layer charge density, which influences the stability and anion exchange capacity of LDHs; for instance, common compositions like Mg/Al LDHs with x \approx 0.33 (corresponding to a Mg:Al ratio of 2:1) exhibit high stability due to optimal charge balancing, while deviations outside 0.2–0.33 can lead to phase instability or segregation of metal hydroxides.[1] Similarly, Zn/Al LDHs with x = 0.25 (Zn:Al = 3:1) are widely used for their robust structure in catalytic applications.[10] A notable variation is the rare Li/Al LDH series, which features a distinct formula [ \mathrm{LiAl}_2 (\mathrm{OH})_6 ]^{+} [ \mathrm{A}^{n-}_{1/n} \cdot m \mathrm{H}_2\mathrm{O} ]^{-}, where the monovalent \mathrm{Li}^{+} occupies octahedral sites alongside \mathrm{Al}^{3+}, creating a fixed 1:2 Li:Al ratio and unique octahedral vacancies that enhance its ion-sieving properties.[11]This compositional flexibility originates from the brucite (\mathrm{Mg(OH)}_2) structure, where partial replacement of \mathrm{Mg}^{2+} by \mathrm{M}^{3+} generates the net positive charge, enabling the incorporation of a wide array of anions to maintain electroneutrality—a feature first systematically explored in synthetic hydrotalcites in the mid-20th century.[12]
Layered Architecture
The layered architecture of layered double hydroxides (LDHs) consists of brucite-like sheets formed by edge-sharing metal hydroxide octahedra. In this structure, each octahedron features a metal cation at its center, coordinated to six hydroxide ions, yielding basic units of [\mathrm{M(OH)_6}], where M represents divalent (e.g., Mg²⁺, Zn²⁺) or trivalent (e.g., Al³⁺, Fe³⁺) cations. These octahedra share edges to create infinite two-dimensional layers with a composition akin to brucite, Mg(OH)₂, but modified by partial substitution of divalent cations with trivalent ones.[13][14]This substitution imparts a net positive charge to the layers, described by the formula [\mathrm{M^{2+}_{1-x} M^{3+}_x (OH)_2}]^{x+}, where x denotes the fraction of trivalent cations, typically ranging from 0.20 to 0.33 to maintain structural stability. The positive layer charge requires balancing by anions located in the interlayer space, though the specific nature of these anions does not alter the intrinsic layer geometry. This charged brucite-like framework distinguishes LDHs from neutral brucite, enabling tunable properties based on metal composition.[13]The positively charged layers stack along the c-axis in ordered sequences that define the polytypes of LDHs. The most common arrangements are the rhombohedral ABC stacking (3R₁ polytype, space group R\bar{3}m) and the hexagonal AB stacking (2H₁ polytype, space group P6₃/mmc), with the 3R₁ being predominant in synthetic Mg-Al LDHs. In carbonate-intercalated variants, the basal spacing d_{003}—the repeat distance perpendicular to the layers—is characteristically 7.8 Å, reflecting the thickness of one layer plus the interlayer region. Variations in stacking can arise from interlayer interactions, but the core octahedral layer motif remains consistent.[14]The parameter x also governs layer curvature and potential structural deviations; at values exceeding 0.33, the elevated charge density from increased trivalent cation incorporation can induce layer bending, defects such as stacking faults, or phase instability. Such extremes often compromise phase purity, limiting stable LDH formation to lower x ranges in most systems.[14]
Interlayer Region
The interlayer region of layered double hydroxides (LDHs) consists of a gallery space between positively charged metal hydroxide layers, occupied by exchangeable anions and water molecules that maintain charge neutrality and structural stability. Common intercalated anions include carbonate (CO₃²⁻), nitrate (NO₃⁻), chloride (Cl⁻), and sulfate (SO₄²⁻), which balance the positive layer charge through electrostatic interactions. The orientation of these anions within the interlayer varies with their size and shape: smaller monovalent anions like Cl⁻ and NO₃⁻ typically adopt a perpendicular or tilted orientation relative to the layers, while larger polyatomic divalent anions such as CO₃²⁻ lie parallel to the layers, influencing the overall packing and accessibility of the gallery.[15][16]Hydration in the interlayer region is characterized by 1 to 4 water molecules per formula unit, forming structured networks that solvate anions and expand the gallery height. At low hydration levels (e.g., 1 H₂O), water molecules arrange in a flat monolayer, whereas higher levels (3–4 H₂O) lead to disordered double layers and significant basal spacing increases, such as expansions up to 0.3–0.5 nm for Cl⁻-intercalated LDHs. This hydration modulates interlayer dynamics, with water-anion hydrogen bonding strengthening as hydration rises, thereby affecting anion mobility and material swelling. For instance, in Mg₄Al₂-LDH systems, NO₃⁻-intercalated variants exhibit stable spacings of approximately 0.89 nm at 3 H₂O per unit cell, compared to 0.73 nm for CO₃²⁻ at lower hydration.[15][17][17]Anion selectivity in the interlayer arises from differences in charge density and ionic size, with LDHs exhibiting a strong preference for divalent anions over monovalent ones due to enhanced electrostatic binding. High-charge-density anions like CO₃²⁻ are favored over NO₃⁻, as evidenced by exchange affinities where CO₃²⁻ intercalation drives forces exceeding those of monovalent counterparts in MgAl-LDHs. Smaller anions are generally preferred for easier gallery insertion, though steric factors can limit larger divalent species like SO₄²⁻ in highly charged systems. This selectivity underpins the tunable nature of LDH interlayers for targeted ion hosting.[18][18][18]The interlayer can accommodate larger organic anions, such as terephthalate, which expand the gallery height significantly—up to approximately 20 Å—through vertical or interdigitated orientations that pillar the structure. These organic intercalates, including terephthalate in Mg-Al LDHs, increase spacing from typical 0.78 nm (e.g., in carbonate forms) to 2.0–2.2 nm, enabling enhanced hosting capacity and altered hydrophobicity compared to inorganic counterparts. Such modifications highlight the versatility of the interlayer for advanced material designs.[19][16]
Synthesis and Preparation
Coprecipitation Method
The coprecipitation method represents the most widely adopted laboratory approach for synthesizing layered double hydroxides (LDHs), relying on the simultaneous precipitation of divalent and trivalent metal hydroxides from aqueous salt solutions under basic conditions.[20] This technique enables the formation of the characteristic brucite-like layers with tunable interlayer anions, such as carbonate.[21]The process begins with the preparation of a mixed metal saltsolution, typically using nitrates or chlorides of metals like magnesium and aluminum (e.g., Mg(NO₃)₂·6H₂O and Al(NO₃)₃·9H₂O) at a M²⁺/M³⁺ molar ratio of 2 to 4, dissolved in deionized water to concentrations of 0.1–2 M.[22] This solution is added dropwise or via peristaltic pump, simultaneously with a basic solution of NaOH (1–3 M) and often Na₂CO₃ (0.5–1 M for carbonate intercalation), to a reactorvessel under vigorous stirring (e.g., 300–1000 rpm) and inert atmosphere (N₂) to minimize CO₂ contamination.[21] The pH is maintained constant at 9–11 using an automated titrator to ensure uniform supersaturation and prevent phase separation into individual metal hydroxides; temperatures range from room temperature to 80°C during precipitation to control nucleation rates.[23] The resulting slurry undergoes aging for 18–72 hours at 25–65°C with gentle stirring, promoting crystal growth and structural ordering without high pressure.[21] The precipitate is then centrifuged, washed repeatedly with deionized water until the pH reaches ~10, and dried at 60–100°C.[22] Key to success is the slow addition rate to manage supersaturation, as rapid mixing can yield amorphous products or impurities.[23]This method's advantages include its simplicity, low equipment requirements, scalability for industrial production, and flexibility in adjusting metal ratios and anions to tailor LDH composition and properties, often yielding crystalline materials with high purity when pH is precisely controlled.[20] Disadvantages encompass the risk of co-precipitated salt impurities (e.g., Na⁺ or NO₃⁻) requiring extensive washing, and potential variability in crystallinity if supersaturation is not optimized, sometimes necessitating post-treatment.[23] Seminal developments trace to early works like Miyata (1975), who demonstrated coprecipitation for hydrotalcite-like compounds using Mg/Al salts at basic pH.A representative example is the synthesis of Mg/Al LDH with CO₃²⁻ intercalation, where a solution of MgCl₂·6H₂O (1 mol) and AlCl₃·6H₂O (0.5 mol) in 700 mL water (Mg/Al = 2) is coprecipitated with NaOH (3.5 mol) and Na₂CO₃ (0.94 mol) at room temperature and pH 10, followed by 18 hours of aging under stirring, yielding a phase-pure product suitable for anion exchange applications.[21]
Hydrothermal and Other Methods
Hydrothermal synthesis of layered double hydroxides (LDHs) involves reacting metal salts in a sealed autoclave under elevated temperatures and pressures, typically ranging from 100 to 200°C for 12 to 48 hours, which promotes the formation of highly crystalline structures with controlled anion incorporation.[24] This method utilizes urea hydrolysis to generate uniform nucleation sites, enabling precise tuning of interlayer anions and enhanced purity compared to ambient conditions.[24] For instance, Zn/Al LDHs synthesized hydrothermally at 100°C for 24 hours have been used to fabricate oriented films on alumina substrates, yielding bilayer membranes with improved structural integrity.Other synthesis approaches expand the versatility of LDH preparation. Ion-exchange methods start with pre-formed LDHs, typically nitrate-intercalated, and involve stirring in excess target anions (e.g., carbonate or phosphate) at pH ≥ 4 under inert atmosphere to achieve selective intercalation without disrupting the layered framework.[24] Calcination-rehydration leverages the structural memory effect: LDHs are heated to 450°C to form mixed oxides, then rehydrated in anion-containing solutions to reconstruct the hydroxide layers, allowing incorporation of complex anions like biomolecules.[24] Sol-gel techniques dissolve metal precursors in solvents, followed by hydrolysis and condensation at 80–150°C, producing nanoscale particles or thin films with high surface areas suitable for coatings.[24] Urea-based homogeneous precipitation heats metal salts with urea at around 90–190°C, where urea decomposes to release CO₂ and NH₃, gradually raising pH for uniform crystal growth and eco-friendly anion control.[24]These methods offer distinct advantages over basic precipitation, such as superior crystallinity, larger particle sizes (up to micrometers), and better morphological control, which are critical for applications requiring mechanical stability or specific orientations.[24] Hydrothermal routes, in particular, yield higher-purity LDHs with tunable interlayer spacing, as demonstrated in Mg/Al systems where 190°C treatment for 1 hour produces carbonate-intercalated variants with enhanced thermal stability.[24]Emerging techniques as of 2025 further optimize synthesis efficiency. Microwave-assisted hydrothermal processes accelerate reactions dramatically; for example, Ni/Co LDHs can be prepared in just 210 seconds at 600 W, forming porous nanospheres with ethylene glycol intercalation and specific capacitances exceeding 2000 F/g, far surpassing conventional hours-long heating.[25] Template-directed approaches, such as oriented self-assembly using structure-directing agents like ethanolamine, enable the fabrication of ultrathin 2D-on-3D hierarchical Co/Ni LDH nanoflowers at ambient temperature, achieving surface areas around 154 m²/g and improved ion accessibility.[26] Recent 2025 advancements include an industrially scalable synthesis of NiFe-LDH at room temperature and atmospheric pressure, facilitating efficient electrocatalysis production,[27] and advanced morphological control via electrochemical deposition and microwave-assisted methods for energy storage applications.[28] These innovations reduce energy consumption and enable scalable production of specialized LDH architectures.[24]
Physical and Chemical Properties
Structural Properties
Layered double hydroxides (LDHs) exhibit a highly ordered layered structure that is primarily confirmed through X-ray diffraction (XRD) analysis, which reveals characteristic basal reflections corresponding to the (00l) planes. For carbonate-intercalated LDHs, the (003) reflection typically appears at approximately 11.5° 2θ, yielding a basal spacing (d003) of about 7.7–7.8 Å, indicative of the interlayer distance accommodating carbonate anions and water molecules.[29] This d-spacing can be calculated using Bragg's law, n\lambda = 2d \sin\theta, where \lambda is the X-ray wavelength (1.54 Å for Cu Kα radiation), confirming the rhombohedral symmetry and layered architecture. Higher-order reflections, such as (006) at around 23° 2θ (d ≈ 3.8 Å), further validate the crystallinity and stacking regularity.[30]Microscopic techniques provide insights into the morphology and surface features of LDHs, typically revealing a platelet-like structure with nanoscale thickness and micrometer-scale lateral dimensions. Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) images show hexagonal or irregular platelets with thicknesses ranging from 5–100 nm and lateral sizes of 0.1–2 μm, depending on synthesis conditions and composition.[31]Atomic force microscopy (AFM) complements these observations by mapping surface topography, highlighting the flat, layered surfaces with roughness on the order of 1–5 nm and occasional stacking irregularities.[32]Spectroscopic methods offer detailed information on the chemical environments within LDHs. Fourier-transform infrared (FTIR) spectroscopy identifies key vibrational modes, including a broad OH stretching band at 3400–3600 cm−1 attributed to hydroxyl groups in the brucite-like layers and interlayer water, alongside sharper peaks around 3600–3700 cm−1 for free OH. Anion modes, such as carbonate asymmetric stretching at 1350–1400 cm−1 and 1500–1600 cm−1, confirm the intercalation and coordination of interlayer species.[33] Solid-state nuclear magnetic resonance (NMR) spectroscopy elucidates the local metal environments; for instance, 27Al NMR reveals octahedral Al3+ sites at chemical shifts of 5–15 ppm in Mg-Al LDHs, while 25Mg NMR distinguishes divalent cation coordination, providing evidence of cation ordering and defects.[34]Structural variability in LDHs arises from modifications and imperfections, affecting the interlayer spacing and overall crystallinity. Pillaring with organic anions, such as carboxylates or polymers, expands the basal spacing to 15–25 Å or more, as the larger guests prop up the layers and enable tunable gallery heights for hosting molecules.[35] Defects, including stacking faults, are common in LDHs with non-ideal x-values (where x = M3+/(M2+ + M3+) deviating from 1/3), leading to broadened XRD peaks and turbostratic disorder observable via simulated diffraction patterns or NMR line broadening.[36]
Thermal and Ion-Exchange Properties
Layered double hydroxides (LDHs) exhibit characteristic thermal decomposition behavior observed through thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC), typically proceeding in three distinct stages. The initial stage involves the loss of interlayer and physisorbed water molecules, occurring between approximately 100–200°C, resulting in a weight loss of about 10–15%. This is followed by dehydroxylation of the brucite-like layers in the 200–450°C range, with a weight loss of around 20%, accompanied by endothermic peaks in DSC profiles. The final stage, above 450°C up to 800°C, entails the decomposition of interlayer anions and structural collapse to form mixed metal oxides, leading to an additional 20–25% weight loss and exothermic events in DSC.A notable feature of LDHs is the "memory effect," whereby calcined LDHs, transformed into layered double oxides (LDOs) at moderate temperatures (300–800°C), can reconstruct their original layered structure upon rehydration in aqueous solutions. This reconstruction facilitates the insertion of new anions into the interlayer region, enabling tailored modifications for various applications. The process relies on the high surface area and reactivity of LDOs, which promote rapid reformation of the hydroxide layers under hydrous conditions.[37]LDHs possess significant anion-exchange capacity (AEC), typically ranging from 2 to 4 meq/g, allowing for the replacement of interlayer anions with external ones in solution. Exchange rates are generally faster for monovalent anions due to their smaller size and lower hydration energy compared to divalent species. Anion selectivity follows the series CO_3^{2-} > SO_4^{2-} > NO_3^- > Cl^-, influenced by factors such as anion charge density, size, and hydration, with carbonate exhibiting the highest affinity owing to its strong electrostatic interactions with the positively charged layers.[38][39]The stability of LDHs is highly pH-dependent, with robust structural integrity maintained in basic conditions (pH > 7), where the layered architecture remains intact. In acidic environments (pH < 7), however, LDHs undergo dissolution, releasing constituent metal cations into solution due to protonation and breakdown of the hydroxide layers. This pH-responsive behavior underpins controlled release mechanisms in anion-exchange processes.[40]
Applications
Environmental and Catalytic Applications
Layered double hydroxides (LDHs) have emerged as effective adsorbents for anion removal in environmental remediation, particularly for pollutants such as phosphates, arsenate, dyes, and hexavalent chromium (Cr(VI)). For instance, Zn1.5Al-LDH exhibits a phosphate sorption capacity of up to 22 mg P/g, achieving over 90% removal efficiency under optimized conditions. Similarly, Mg/Al-LDH composites demonstrate high efficiency in removing arsenate and Cr(VI) from aqueous solutions, with removal rates exceeding 90% for Cr(VI) at pH 3–7 due to the material's layered structure facilitating anion uptake. In dye removal applications, LDH-based materials, including those derived from halloysite-Mg/Al composites, achieve efficient adsorption of organic dyes through surface interactions, often surpassing 95% removal for common textile effluents. These adsorption processes leverage the high anion-exchange capacity of LDHs, typically ranging from 200–400 meq/100 g, enabling selective binding in contaminated water systems.The primary mechanisms for anion removal by LDHs involve anion exchange, where interlayer anions are replaced by target pollutants, and surface complexation, where pollutants form inner-sphere complexes with metal hydroxide layers. For Cr(VI), this includes both electrostatic attraction and redox processes, with Ca4Al2-Cl LDH showing an affinity increase to 3.2 × 10^5 mL/g via Cr-Ca coordination and nano-confinement effects. Phosphate and arsenate removal follows similar pathways, with intercalation being more efficient than precipitation, as demonstrated by Mg/Al-LDHs achieving higher capacities at neutral pH. These mechanisms ensure LDHs' reusability, with regeneration via simple anion desorption maintaining over 80% efficiency after multiple cycles.In wastewater treatment, LDH-based membranes have been developed for heavy metal removal, combining adsorption with filtration to achieve selective rejection rates above 95% for ions like Pb(II) and Cd(II). Recent advancements include LDH-cellulose hybrid membranes that enhance flux while targeting pharmaceuticals, such as antibiotics in industrial effluents. For example, ZnAl-LDH and CoZr-LDH adsorbents remove over 90% of metronidazole from pharmaceutical wastewater through π-π interactions and hydrogen bonding, with 2025 studies highlighting their scalability for real-world antibiotic pollution control.LDHs also serve as versatile catalysts in environmental processes, often acting as precursors to mixed metal oxides after calcination, which retain high surface areas (up to 200 m²/g) and basic sites for reactions like CO2 methanation and biodiesel production. In CO2 methanation, Ni-based LDHs with Mg or Ca yield dual-function materials that capture and convert CO2 to methane with conversions exceeding 90% at 300–400°C, attributed to strong metal-support interactions. For biodiesel transesterification, calcined Mg-Al and Zn-Al LDHs catalyze the reaction of triglycerides with methanol, achieving yields over 95% under mild conditions (60–100°C), with hydrothermal reconstruction enhancing basicity and reusability up to 10 cycles.Pillared LDHs, where interlayer spacing is expanded with polyoxometalates or metal oxides, excel in volatile organic compound (VOC) oxidation, enabling complete conversion of toluene or ethanol at temperatures below 300°C. Cu-Cr LDH-derived oxides, for instance, show 100% VOC removal efficiency due to synergistic redox sites, outperforming non-LDH catalysts in stability and low-temperature activity.In photocatalysis, LDH-TiO2 composites facilitate dye degradation under UV or visible light, with materials like MgAl-TiO2 achieving over 90% removal of methylene blue in 2 hours via enhanced charge separation. Recent 2021–2025 developments focus on hydrogen production, where NiAl-LDH-modified systems yield rates exceeding 1000 μmol/g/h under solar irradiation, as seen in trimetallic Ni-Co-Al LDHs coupled with g-C3N4, promoting water splitting through improved electron transfer and band alignment. These composites also enable simultaneous H2 evolution and pollutant degradation, with NiAl-LDH-TiO2 hybrids reaching 1063 μmol/g/h for H2 while degrading dyes.
Biomedical and Pharmaceutical Applications
Layered double hydroxides (LDHs) have emerged as promising nanomaterials in biomedical and pharmaceutical applications due to their biocompatibility, tunable layered structure for hosting bioactive molecules, and ability to enable controlled release in physiological environments. Their anionic clay-like composition allows for the intercalation of drugs, nucleic acids, and imaging agents, facilitating targeted delivery while minimizing off-target effects. High biocompatibility is evidenced by low cytotoxicity in various cell lines, with LDHs degrading into biocompatible ions like Mg²⁺ and Al³⁺ that support cellular processes without inducing inflammation in vivo.[41][42]In drug delivery, LDHs excel through anion exchange, enabling high loading of non-steroidal anti-inflammatory drugs (NSAIDs) such as ibuprofen. For instance, MgAl-LDHs intercalate ibuprofen with capacities up to 330 mg/g, expanding interlayer spacing to accommodate the guest molecules. This results in sustained release profiles, with up to 100% ibuprofen liberation over 100 minutes at neutral pH via anion exchange, extendable to hours in simulated physiological conditions. The pH-responsive nature of LDHs further enhances efficacy, as dissolution accelerates in acidic tumor microenvironments (pH ~5-6), promoting targeted release and reducing systemic exposure.[43][44]LDHs also serve as non-viral vectors in gene therapy by complexing with DNA or RNA for efficient transfection while mitigating toxicity associated with viral methods. MgAl-LDH nanoparticles form stable complexes with siRNA or miRNA at mass ratios of 10-50, achieving encapsulation efficiencies of 73-76% and cellular uptake exceeding 90% in mesenchymal stem cells at concentrations of 20-100 nM. Transfection success includes up to 80% gene knockdown for siRNA delivery, with low cytotoxicity at optimal ratios (e.g., no impact on cell viability at mass ratio 10), avoiding immunogenicity and genomic integration risks of viral vectors. Applications span cancer silencing and regenerative therapies, such as miRNA-16 overexpression for bone repair in scaffold-integrated systems.[42][45]For imaging and combined therapy, LDH nanoparticles function as contrast agents and carriers, leveraging their paramagnetic properties and responsiveness. Manganese-based LDHs (Mn-LDHs) act as T1-weighted MRI contrast agents with ultrasensitive pH response, exhibiting longitudinal relaxivity of 9.48 mM⁻¹s⁻¹ at pH 5.0 (tumor-like acidity) versus 1.16 mM⁻¹s⁻¹ at pH 7.4, enabling clear in vivo tumor visualization in mice for up to 2 days. These nanoparticles also support photothermal therapy through near-infrared absorption, enhancing theranostic potential. In cancer treatment, doxorubicin (DOX)-loaded MgAl-LDHs demonstrate pH-triggered release in acidic environments, achieving significant tumor growth inhibition in vivo models with improved cellular uptake and reduced cardiotoxicity compared to free DOX, as shown in 2018 studies with ongoing advancements in targeted nanomedicine by 2025.[46][47][41]In tissue engineering, LDHs form bioactive scaffolds for bone regeneration, promoting osteogenesis and vascularization due to ion release and structural mimicry of bone minerals. Composites like hydroxyapatite/MgAlEu-LDH exhibit 3.18-fold increased bone formation in vivo via Wnt/β-catenin pathway activation and M2 macrophage polarization, with biocompatibility confirmed by >95% cell viability at 500 µg/mL in osteoblast cultures and no inflammation after 28-day implantation in rats. Low cytotoxicity (e.g., >90% viability in PC12 cells at 50 µg/mL for MgFe-LDH) supports their use in 3D-printed or hydrogel scaffolds, such as PCL/gelatin-MgAl-LDH, which enhance mechanical strength and osteogenic differentiation, with 2025 advances including personalized Cu-doped LDH scaffolds for precision regeneration.[41]
Energy and Materials Applications
Layered double hydroxides (LDHs) play a significant role in energy storage and conversion technologies, leveraging their brucite-like layered structure for efficient ion diffusion, redox activity, and high surface area. In supercapacitors, Ni- and Co-based LDHs function as pseudocapacitive electrodes, delivering exceptional specific capacitances often exceeding 1000 F/g due to faradaic reactions involving metal hydroxide/oxyhydroxide transitions. For example, thin-layer NiCo-LDH nanosheets exhibit a specific capacitance of 3982.5 F/g at a current density of 1 A/g, with excellent rate capability and cycling stability retaining over 93% after 5000 cycles. Recent progress from 2024–2025 emphasizes flexible supercapacitors fabricated via exfoliation of LDHs, such as NiCr-LDH nanosheets on exfoliated nano-graphite flakes, which enhance mechanical flexibility, energy density, and suitability for wearable electronics while maintaining high capacitance and long-term durability.[48][49]LDHs also serve as anode materials in rechargeable batteries, capitalizing on their interlayer expandability for accommodating alkali metal ions during charge-discharge cycles. In lithium-ion batteries, NiAl-LDH electrodes demonstrate high initial discharge capacities of up to 2586 mAh/g at 0.05 A/g, attributed to reversible Li⁺ intercalation and structural stability that mitigates volume expansion. Similarly, in sodium-ion batteries, MnCo-LDH hybrids with graphene deliver specific capacities of 941 mAh/g after 200 cycles at 0.2 A/g, benefiting from enhanced conductivity and ion transport pathways. Beyond storage, Pt-free NiFe LDHs excel as catalysts for the hydrogen evolution reaction (HER) in alkaline electrolytes, achieving low overpotentials below 100 mV at 10 mA/cm²—for instance, 89 mV for NiFeAu-LDH—due to optimized electronic structure and active edge sites that facilitate water dissociation and hydrogen adsorption.[50][51][52]In advanced materials applications, LDHs enhance the performance of polymer composites for fire safety and corrosion protection. MgAl-LDH incorporated into polymers like polystyrene acts as an eco-friendly flame retardant, reducing peak heat release rates by nearly 50% through intumescentchar formation, endothermic dehydration, and release of water/CO₂ that dilutes combustible gases. For anticorrosion, LDH-based coatings on metals such as Al and Mg alloys provide active protection via anion-exchange mechanisms, where aggressive ions like Cl⁻ are trapped and replaced by corrosion inhibitors, leading to self-healing and impedance increases over 10⁶ Ω·cm² after prolonged exposure. These properties stem from the tunable interlayer anions and high anion-exchange capacity of LDHs.[53][54]Emerging developments integrate LDHs with perovskites in hybrid structures for photovoltaic devices, improving charge extraction and stability at interfaces. ZnAl-LDH precursors as scaffolds in CH₃NH₃PbI₃ perovskite solar cells yield power conversion efficiencies up to 18.54%, surpassing control devices by reducing recombination and enhancing hole transport, with retained performance under ambient conditions. Such hybrids, explored in recent studies up to 2025, highlight LDHs' potential to push efficiencies beyond 20% by addressing hysteresis and degradation in next-generation solar technologies.[55]
Natural Minerals and Occurrence
Hydrotalcite Supergroup
The Hydrotalcite supergroup encompasses natural layered double hydroxides (LDHs) officially recognized by the International Mineralogical Association (IMA), with more than 50 approved species as of 2025.[56][57] Since 2012, the supergroup has expanded with over 10 new species, including marioantofilliite (IMA 2025-012).[58] These minerals are classified into eight groups based on the M^{2+}:M^{3+} octahedral cation ratio (typically 2:1 or 3:1) and interlayer anion composition, which determine the layer stacking and basal spacing; polytypic variations within groups include 2-layer (e.g., 2H) and 3-layer (e.g., 3R) structures.[56] The supergroup's nomenclature was formalized in 2012, establishing a stable taxonomic framework that emphasizes structural and compositional consistency across species.[56]Common minerals in the supergroup belong primarily to the hydrotalcite group (3:1 ratio, ~7.8 Å spacing) and quintinite group (2:1 ratio, ~7.8 Å spacing). Representative examples include hydrotalcite, with ideal formula \mathrm{Mg_6Al_2(OH)_{16}(CO_3)\cdot 4H_2O}, takovite, \mathrm{Ni_6Al_2(OH)_{16}(CO_3)\cdot 4H_2O}, and pyroaurite, \mathrm{Mg_6Fe^{3+}_2(OH)_{16}(CO_3)\cdot 4H_2O}.[56] These species highlight typical carbonate-intercalated structures, where the brucite-like octahedral layers accommodate divalent and trivalent cations in a ordered arrangement.[56]Rarer species occur in groups with distinct interlayer anions or cation combinations, such as the woodwardite group (~8.9 Å spacing, sulfate-based) and hydrocalumite group (unique Ca-Al framework). Woodwardite features Cu-Zn-Al sulfate compositions, exemplified by (\mathrm{Cu,Zn})_{1-x}\mathrm{Al_x(OH)_2(SO_4)_{x/2}\cdot nH_2O, while hydrocalumite represents Ca-Al chloride variants, such as \mathrm{Ca_2Al(OH)_6Cl\cdot nH_2O}.[56]Cation substitutions in the supergroup are extensive within the octahedral sites, with divalent M^{2+} typically including Mg, Fe^{2+}, Ni, Zn, and Ca, and trivalent M^{3+} comprising Al, Fe^{3+}, and Cr.[56] These variations allow for diverse end-member compositions while maintaining the overall positively charged layer structure that requires anion balancing.[56]Interlayer anions are predominantly carbonate (CO_3^{2-}), which stabilizes the 7.8 Å spacing in most common species, but chloride (Cl^-) and sulfate (SO_4^{2-}) anions are also significant, leading to expanded basal spacings in their respective groups.[56]All species exhibit layered structures with hexagonal or rhombohedral symmetry (space groups such as R\bar{3}m or P6_3/mmc), and basal spacings ranging from 7.8 Å (carbonate-dominated) to 11 Å (highly hydrated or larger-anion interlayers).[56] This structural uniformity underpins the supergroup's definition, with polytypes arising from different stacking sequences of the hydroxide layers.[56]
Geological Formation
Layered double hydroxides (LDHs) primarily form in natural geological settings through low-temperature precipitation, typically below 200°C, in alkaline environments such as those associated with serpentinized ultramafic rocks or evaporitic deposits. During the carbonation of mafic and ultramafic rocks, alkaline fluids rich in Mg²⁺ and other divalent cations interact with CO₂-bearing waters, leading to the precipitation of LDH phases like hydrotalcite, where trivalent cations such as Al³⁺ substitute into brucite-like layers, balanced by interlayer anions like CO₃²⁻.[59] This process is prominent in ophiolite complexes, where hyperalkaline springs from serpentinized peridotites in regions like Oman and the Ligurian Alps, Italy, deposit LDH minerals alongside brucite.[59] In evaporitic contexts, LDHs such as dritsite form via diagenesis in halite-carnallite sequences, driven by the concentration of Li⁺, Al³⁺, and Cl⁻ in saline pore waters.[60]Secondary alteration contributes significantly to LDH formation, particularly in weathering zones where precursor minerals like serpentine or clays undergo Mg leaching and Al enrichment under oxidative or hydrolytic conditions. In hydromorphic soils and sediments, partial oxidation of Fe(II)-bearing phases in anoxic, waterlogged environments promotes the transformation into LDHs such as fougèrite, a green rust variant stabilized by interlayer carbonates and varying Fe²⁺/Fe³⁺ ratios.[61] This mechanism is evident in gleysols, where bacterial Fe reduction and fluctuating redox potentials facilitate LDH persistence.[62]LDH minerals occur worldwide in diverse settings, including sediments and soils; for instance, hydrotalcite is documented in the Caselton mine, Nevada, USA, within oxidized manganese deposits, and stichtite in lateritized komatiites at Dundas, Tasmania, Australia.[63]Fougèrite appears in forested gleysols near Fougères, Brittany, France, and related phases in maritime marshes like Mont Saint-Michel Bay.[61] These occurrences are rarer in volcanic terrains but noted in altered basaltic glasses. LDHs exhibit stability in pH ranges of 8–12 within CO₂-rich aqueous systems, resisting dissolution under mildly alkaline to neutral conditions.[64] Recent 2025 investigations have linked natural LDH formation to carbon sequestration in submarine environments, such as altered oceanic crust near deep-sea vents, where low-temperature carbonation traps CO₂ in interlayer anions during serpentinization.[65]