Fact-checked by Grok 2 weeks ago

Anhydrite

Anhydrite is a composed of with the CaSO₄. It crystallizes in the orthorhombic system, typically forming massive, granular, or fibrous aggregates that are colorless, white, or pale shades of blue, gray, or violet. Unlike its hydrated counterpart (CaSO₄·2H₂O), anhydrite lacks molecules in its and readily converts to gypsum upon exposure to under surface conditions. Anhydrite commonly occurs in sedimentary deposits, often as a major component in sequences formed by the of waters, and in cap rocks overlying domes, where it arises from the of gypsum at elevated temperatures or depths. Its hardness ranges from 3 to 3.5 on the , with a specific gravity of about 2.9 to 3.0, making it denser than gypsum. Industrially, anhydrite serves as a source of sulfate in production and as a filler in various applications due to its and moisture resistance.

Mineralogical Characteristics

Chemical Composition and Polymorphs

Anhydrite is the anhydrous calcium sulfate mineral with the idealized chemical formula CaSO₄, consisting of calcium, sulfur, and oxygen in a 1:1:4 molar ratio, corresponding to 29.43% Ca, 23.55% S, and 46.99% O by weight, or equivalently 41.19% CaO and 58.81% SO₃. Natural specimens may contain minor impurities such as Sr, Na, or Mg substituting for Ca, or silica inclusions, but these do not alter the primary composition. The anhydrite specifically refers to the orthorhombic polymorph of CaSO₄, denoted β-anhydrite, which is the thermodynamically stable form under conditions at surface temperatures and pressures. This polymorph features a structure with distorted SO₄ tetrahedra linked by Ca atoms in a pseudo-cubic , yielding perfect on {010}, {100}, and {001}. A metastable hexagonal polymorph, γ-anhydrite (or soluble anhydrite), forms primarily through rapid dehydration of between 75–105°C and exhibits higher in ; it transforms reconstructively to β-anhydrite upon heating to 150–900°C, often producing oriented triplet crystals. At temperatures exceeding 1200°C, a high-temperature cubic polymorph, α-anhydrite, emerges via polymorphic inversion from β-anhydrite, reverting to the orthorhombic form upon cooling; this phase is not observed in natural geological settings due to its thermal instability.

Crystal Structure

Anhydrite, chemical formula , crystallizes in the with Amma (No. 74). The unit cell parameters are a = 6.993(2) , b = 6.995(2) , c = 6.245(1) , and Z = 4. These dimensions reflect a nearly square cross-section in the a-b plane, contributing to the mineral's prismatic or tabular habits. The atomic arrangement features isolated SO₄ tetrahedra, where sulfur is centrally bonded to four oxygen atoms with S-O distances typically ranging from 1.47 to 1.49 . Calcium cations occupy sites coordinated to eight oxygen atoms, forming distorted dodecahedra or irregular polyhedra with Ca-O bond lengths varying between approximately 2.41 and 2.85 . These Ca polyhedra link the sulfate tetrahedra into a three-dimensional via corner- and edge-sharing, stabilizing the anhydrous structure without present in hydrated polymorphs like . Crystal habits include equant forms with large pinacoidal faces or thick tabular parallel to {010}, {100}, or {001}; elongation along , , or is common, with crystals up to 15 cm reported and over 40 distinct forms documented. Twinning occurs on {101} or {111}, though less frequently than in . The structure's refinement, initially determined in 1925 and later improved through , confirms its of about 2.96 g/cm³, consistent with measured values.

Physical and Optical Properties

Anhydrite crystallizes in the orthorhombic system, typically forming granular, nodular, fibrous, or massive aggregates, with rarer prismatic or tabular crystals. It exhibits perfect on {010}, nearly perfect on {100}, and good to imperfect on {001}, often yielding cubic-like fragments due to the orthogonal directions. The mineral is brittle, with a Mohs of 3 to 3.5 and a of 2.98 g/cm³. Common colors include colorless, , gray, blue, , pink, or brown, with a white to grayish streak and vitreous to pearly luster. Anhydrite is transparent to translucent, displaying uneven to splintery . Optically, anhydrite is biaxial positive, with refractive indices of nα = 1.567–1.574, nβ = 1.574–1.579, and nγ = 1.609–1.618, resulting in a birefringence of δ = 0.042–0.044. The 2V measures 36° to 45°, and varieties exhibit visible pleochroism with absorption Z > Y > X. It shows strong (r < v) and moderate relief in thin section, with polysynthetic twinning observable under crossed polars.

Formation Processes

Evaporitic Precipitation

Evaporitic precipitation of anhydrite occurs primarily in restricted marine basins under arid climatic conditions, where progressive evaporation of seawater concentrates dissolved ions until calcium sulfate reaches supersaturation. This process typically follows the deposition of early carbonate evaporites, such as calcite and dolomite, when the original seawater volume has been reduced by approximately 85-90% through evaporation, as demonstrated in classic experiments by Usiglio in 1884. At this stage, calcium sulfate precipitates, initially favoring gypsum (CaSO₄·2H₂O) in near-surface, cooler brines due to its higher solubility at ambient temperatures. Direct precipitation of anhydrite (CaSO₄) from brine becomes dominant in warmer, deeper hypersaline settings or where temperatures exceed the gypsum-anhydrite transition point of approximately 58°C at atmospheric pressure, beyond which anhydrite is the thermodynamically stable phase. In such environments, decreasing solubility of anhydrite with increasing temperature and salinity facilitates its crystallization as bedded, nodular, or enterolithic fabrics, often interbedded with in mature evaporite sequences. These textures, including characteristic "chicken-wire" structures, reflect synsedimentary growth in sabkha-like marginal settings or density-stratified brines, where displacive nodular growth displaces primary sediments. The solubility of calcium sulfate in NaCl-dominated brines, mimicking evaporated seawater, exhibits a retrograde behavior with temperature, promoting anhydrite precipitation over gypsum at elevated thermal conditions common in subtropical to tropical evaporite basins. Empirical models confirm that anhydrite solubility minima occur around 150-200°C in saline solutions, but surface and shallow subsurface precipitation is driven by evaporative concentration rather than deep thermal effects. Major geological examples include the Permian Zechstein Basin in Europe and Miocene evaporites in the Carpathian foreland, where thick anhydrite units overlie gypsum and underlie halite, recording repeated cycles of flooding and desiccation. This precipitation mechanism accounts for the bulk of global anhydrite reserves, underscoring its role as a key indicator of ancient paleoclimate and basin hydrology.

Diagenetic and Hydrothermal Alteration


Diagenetic alteration of commonly involves the dehydration of precursor during burial, transforming CaSO₄·2H₂O into anhydrous CaSO₄ at depths where temperatures exceed 60°C under lithostatic pressures of approximately 500-1000 m, depending on interstitial fluid salinity and composition. This process releases up to 480 kg of water per cubic meter of gypsum, facilitating pressure solution, fluid migration, and secondary porosity development in sequences. Resulting textures include nodular, poikilotopic, or enterolithic forms, often preserving primary sedimentary structures while altering mineral fabric. In environments, early diagenetic anhydrite precipitates supratidally as hair-cream or massive varieties, influenced by groundwater evaporation and organic matter stabilization.
Later diagenetic phases may involve anhydrite calcitization, where sulfate ions exchange with carbonate in dolomitic facies, or replacement by authigenic quartz in dissolution pores, enhancing reservoir quality in carbonate-evaporite systems. Upon uplift and exposure to meteoric or undersaturated waters, anhydrite undergoes rehydration to secondary , forming porphyroblastic or fibro-radiating crystals and contributing to karstification or collapse structures. These transformations are recorded in isotopic signatures, with δ¹⁸O and δ³⁴S values reflecting fluid-rock interactions during burial-exhumation cycles. Hydrothermal alteration of anhydrite occurs in high-temperature fluid systems, such as mid-ocean ridges, where anhydrite precipitates from seawater at temperatures above 150°C and pressures around 500 bar, forming veins, stockworks, or permeability seals that modulate fluid flow and metal transport. In basaltic hosts, surface pillow lavas exhibit anhydrite replacement by chlorite or sulfides via interaction with sulfate-undersaturated fluids, indicating high-temperature seafloor alteration up to 320°C. Precipitation-dissolution cycles of anhydrite in mixing zones enhance subseafloor reservoirs by reducing permeability in reaction zones while promoting focused upflow. In continental arc settings, magmatic-hydrothermal fluids sulfidize anhydrite, converting it to pyrite or chalcopyrite and mobilizing metals through sulfate reduction. Such processes are evident in isotopic and trace element profiles, distinguishing hydrothermal from diagenetic origins.

Microbial and Low-Temperature Mechanisms

Microbial activity can facilitate anhydrite precipitation at temperatures as low as 35°C, challenging the conventional requirement for elevated thermal conditions exceeding 50–60°C. Laboratory experiments using evaporated seawater inoculated with natural microbial communities from hypersaline settings, such as salterns, demonstrated direct formation of microcrystalline anhydrite aggregates through biologically mediated nucleation. These processes likely involve microbial extracellular polymeric substances (EPS) that lower the energy barrier for crystal nucleation or alter local solution chemistry via metabolic byproducts, enabling supersaturation and precipitation under otherwise undersaturated conditions for pure inorganic systems. In arid environments, desert microorganisms, including cyanobacteria and fungi within biological soil crusts, induce gypsum dehydration to anhydrite by extracting structural water from CaSO₄·2H₂O crystals to sustain hydration. This biophysically driven process occurs at ambient surface temperatures (typically 20–40°C), producing nanosized anhydrite crystals as a byproduct; the microbes generate acidic microenvironments (pH ~4–5) via organic acid excretion, which accelerates the phase transition by destabilizing gypsum lattices. Field observations from gypsum dunes in the White Sands National Park, New Mexico, corroborate this, with microbial filaments observed embedded in resulting anhydrite nanostructures. Bacterial strains such as Pseudomonas fluorescens influence calcium sulfate speciation in saline solutions by producing biosurfactants and altering ionic adsorption, promoting anhydrite over gypsum formation even at low salinities and temperatures below 40°C. These effects stem from bacteria-mediated changes in solution interfacial tension and nucleation sites, as evidenced in controlled precipitation assays where microbial presence increased anhydrite yield by up to 30% compared to sterile controls. Non-microbial low-temperature mechanisms remain limited but include advective transport of low-water-activity brines in porous media, where continuous solute influx sustains supersaturation without reliance on evaporation alone. Such conditions, modeled for or terrestrial subsurface flows, allow sporadic anhydrite nucleation at 20–50°C, though rates are orders of magnitude slower than microbial pathways and require specific brine compositions (e.g., Mg/Ca ratios >5). Empirical data from flow-through experiments indicate that dissolution-reprecipitation cycles in undersaturated fluids can yield secondary anhydrite at these temperatures, but primary precipitation is thermodynamically inhibited below ~18°C in dilute systems.

Geological Occurrences

Evaporite Basins and Sequences

Anhydrite precipitates in basins under hypersaline conditions where exceeds inflow, concentrating brines to the point of sulfate , typically following initial deposition. These basins often develop in tectonically restricted settings, such as rifted margins or foreland depressions, leading to thick accumulations of sulfate-dominated strata adjacent to basin margins. In sabkha-like marginal environments, anhydrite forms preferentially over due to elevated temperatures and , resulting in nodular or enterolithic fabrics within the . Evaporite sequences containing anhydrite exhibit cyclic bedding patterns driven by oscillatory sea-level changes or reflux, with upward stratigraphic trends from and anhydrite at the base to and salts at the top. Each cycle often comprises -anhydrite--anhydrite- units, bounded by solution disconformities, as observed in Permian deposits of the Hugoton Embayment, , where such sequences reach thicknesses exceeding 300 meters. In deeper basinal settings, like the Permian Formation of the Delaware Basin, West Texas, finely laminated anhydrite alternates with in barred-basin facies, recording perennial negative . Notable examples include the Middle Badenian evaporites of the Carpathian Foredeep in , featuring marginal sulfate platforms with anhydrite up to hundreds of meters thick transitioning basinward into . In the Southern Permian Basin of , giant polygonal anhydrite ridges, 2–4 km in diameter, mark paleosabkha margins within Zechstein-equivalent sequences. These formations underscore anhydrite's role as a stratigraphic marker for paleoclimate and hydrology, with diagenetic conversion from primary common in buried sections.

Salt Domes and Cap Rocks

Anhydrite commonly forms the basal zone of cap rocks overlying salt domes, which are diapiric structures where buoyant sequences, primarily , intrude overlying sediments due to density contrasts. These domes pierce and strata in regions like the U.S. , where approximately 65% of 329 onshore salt stocks exhibit anhydrite-dominated cap rocks up to 300 meters thick. Cap rock development occurs through episodic dissolution of by undersaturated meteoric or formation waters circulating at the dome crest, leaving behind insoluble residues such as the 1-5% impurities originally disseminated in the salt mass. This residual accumulates, compacts, and recrystallizes into a low-permeability layer, often exhibiting laminated or nodular textures reflective of repeated dissolution-precipitation cycles. Overlying the zone, secondary and zones may develop via and precipitation from reduction or fluid interactions. In Gulf Coast examples, such as the Big Hill dome in , cap rocks span 850-1300 feet thick with a distinct anhydrite base transitioning upward to and , formed over to timescales. Similarly, at Tatum dome in , the 500-600 foot thick cap rock mirrors regional zoning patterns, where anhydrite's content influences associated mineralization like sulfides. These structures enhance hydrocarbon entrapment by providing seals against vertical migration, with anhydrite's low (typically <5%) and anhydrite's chemical stability preventing fluid escape. Additionally, biogenic processes, including microbial reduction, contribute to carbonate caps within anhydrite matrices, altering permeability and hosting elemental sulfur deposits in some domes.

Other Associations and Localities

Anhydrite occurs in hydrothermal vein systems, where it precipitates from sulfate-rich fluids associated with ore mineralization, distinct from primary evaporitic settings. In such environments, it often accompanies metallic deposits like gold and iron ores, forming as a gangue mineral during fluid circulation in fractured host rocks. For instance, at the Lienetz orebody within the on , , anhydrite constitutes prominent vein arrays linked to early porphyry-style magmatic-hydrothermal brecciation and vein formation, with veins exhibiting dynamic structural evolution under deformation. In igneous and metamorphic terrains, anhydrite appears in associations with magnetite deposits, interpreted as syngenetic precipitates contemporaneous with ore formation. A key example is the Lyon Mountain magnetite deposit in the northeastern Adirondacks, New York, where anhydrite occurs intergrown with magnetite and other sulfates, likely derived from contemporaneous sulfate-rich fluids during Grenville-age metamorphism and igneous activity around 1,150–1,000 million years ago. Secondary occurrences include karstic aquifers and altered limestones, where anhydrite persists as relic evaporitic material modified by dissolution and recrystallization. In the Northern Apennines of Italy, the Poiano spring drains a gypsum-anhydrite karst system enriched in NaCl, with anhydrite lenses at depth contributing to the aquifer's saline output through selective dissolution processes. Limited reports also note anhydrite in zeolite-filled amygdules of basaltic flows and volcanic fumaroles, though these are rare and typically subordinate to primary evaporite-derived sources.

Historical and Scientific Development

Discovery and Etymology

Anhydrite was first discovered in 1794 in a salt mine near Hall in Tirol, Austria (now Innsbruck), where it occurred in evaporite deposits alongside gypsum. The mineral was initially described as a distinct species after analysis revealed its anhydrous composition, differing from the hydrated calcium sulfate of gypsum, though early observations may have confused it with other sulfate minerals. Austrian naturalist Nicolaus Poda von Neuhaus reportedly provided one of the earliest accounts, naming a similar material muriacite based on its association with salt (muria), but subsequent chemical verification in the late 18th century confirmed its unique identity. The name "anhydrite" was formally proposed in 1804 by German mineralogist Abraham Gottlob Werner, who emphasized its lack of water of crystallization compared to . Derived from the Greek an- (without) and hydor (water), the term anhydros underscores the mineral's anhydrous formula CaSO₄, reflecting Werner's systematic approach to mineral nomenclature based on compositional differences. This naming aligned with early 19th-century advances in crystallography and chemistry, which distinguished anhydrite's orthorhombic structure and dehydration behavior from related .

Key Advances in Petrogenesis

The prevailing early 20th-century view held that most anhydrite formed secondarily through dehydration of gypsum under burial conditions, driven by elevated temperatures and pressures that expelled water of crystallization. This interpretation aligned with phase stability data indicating anhydrite's thermodynamic preference above approximately 58°C at atmospheric pressure, as determined from solubility experiments by Posnjak in 1938. A pivotal advance came in the 1960s with detailed petrological examinations of ancient evaporites, revealing fabrics such as nodular, enterolithic, and cumulate textures inconsistent with post-depositional deformation of gypsum but indicative of primary or syndiagenetic crystallization. Shearman's 1961 study classified these structures in Paleozoic and Mesozoic sequences, attributing them to displacive growth from interstitial brines rather than compactional alteration. Concurrent observations from modern sabkha environments in the , documented by Shearman (1963) and Butler (1969), demonstrated how anhydrite nodules precipitate via capillary evaporation of seawater-derived brines in supratidal sediments, forming at shallow subsurface depths (typically <1 m) under low-temperature conditions (20–40°C). These analogs established that primary anhydrite could nucleate directly in high-salinity, low-water-activity settings without requiring deep burial. Further progress in the 1970s–1980s involved fluid inclusion and isotopic analyses, which quantified formation temperatures and brine compositions. Homogenization temperatures from primary inclusions in bedded anhydrites ranged 40–80°C, supporting shallow diagenetic origins rather than solely metamorphic overprints, while sulfur and oxygen isotopes traced marine sulfate sources diluted by minor freshwater influxes. Hardie et al. (1985) refined criteria to differentiate primary displacive features from secondary replacements, emphasizing textural preservation in undeformed sequences. Recent experimental and modeling advances have addressed kinetic barriers to low-temperature precipitation, resolving aspects of the gypsum-anhydrite sequence. Laboratory simulations since the 2010s show direct nucleation at 35°C from microbially modified evaporated seawater, where organic films inhibit gypsum overgrowth and promote metastable bassanite intermediates that invert to . These findings, corroborated by reactive transport models, indicate that brine flow dynamics and minor impurities enable primary surface or near-surface formation, challenging earlier burial-centric models and highlighting microbial catalysis in ancient evaporites.

Applications and Utilizations

Industrial Extraction and Processing

Anhydrite is extracted primarily from underground evaporite deposits using room-and-pillar mining techniques, which involve selective drilling and blasting to create chambers while leaving supporting pillars. This method allows access to thick, bedded sequences at depths exceeding 200 meters, as practiced in European operations such as those in Lorraine, France. Major producing regions include Europe (with capacities around 700,000 tons annually across sites in France, Spain, and Germany), the United States (deposits in New York, Michigan, and New Mexico), and Canada (Alberta Triassic beds). Solution mining is less common due to anhydrite's lower solubility compared to associated halite, though it occurs in some salt dome contexts. Post-extraction, raw undergoes crushing and screening to uniform sizes, followed by washing to remove impurities like clay or salt, and classification by particle size. Grinding mills reduce it to fine powders suitable for industrial applications, with optional surface treatments or drying to achieve specific reactivities. In cement production, ground serves as a sulfate regulator, blended into clinker raw mixes at 3-5% to control setting time and prevent flash set; it replaces due to its anhydrous nature, which minimizes water introduction. The historical anhydrite process, used in the UK until the 1970s, integrated it into rotary kilns under reducing conditions to yield cement clinker and sulfur dioxide for sulfuric acid recovery, reducing limestone needs by up to 20%. Synthetic anhydrite, derived as a byproduct from hydrofluoric acid production via fluorspar-sulfuric acid reactions or phosphoric acid plants (phosphogypsum dehydration), supplements natural sources and undergoes similar grinding and activation steps. Activation for binder use involves chemical additives like ground or citric acid to enhance hydration kinetics, enabling self-leveling screeds with improved fluidity over cement-based alternatives. Quality control emphasizes low impurities (e.g., <0.5% silica) to avoid defects in end products like , where excess anhydrite can inhibit strength development.

Construction and Materials Science

Anhydrite, or anhydrous (CaSO₄), finds application in construction as a component in flooring screeds, where it forms the basis of calcium sulfate-based systems that self-level to create smooth, durable subfloors. These anhydrite screeds are pumped in liquid form, allowing rapid coverage of large areas up to 4-5 cm thick, and they dry faster than traditional sand-cement screeds, often becoming walkable within 24 hours and ready for finishes like tiles or wood in 48 hours under controlled humidity below 75%. Their low shrinkage and thermal conductivity make them suitable for underfloor heating installations, reducing cracking risks compared to cement-based alternatives. In cement production and modification, anhydrite acts as a setting regulator, similar to gypsum but with distinct hydration kinetics that influence ettringite formation and early strength development. Studies on alkali-activated Portland cement show that incorporating 5-10% anhydrite enhances hydration rates and achieves peak compressive strengths of up to 50 MPa at 28 days, due to optimized sulfate availability without excessive expansion. In sulfoaluminate cements, anhydrite dosages around 10% promote denser microstructures via accelerated ye'elimite hydration, improving resistance to sulfate attack while maintaining workability. Anhydrite-based cements, produced by calcining natural or waste gypsum (e.g., phosphogypsum) at 800-1000°C to form stable β-anhydrite III, serve as binders in specialized mortars, grouts, and self-leveling compounds, offering lower energy use than Portland cement clinker production. These cements, activated with salts like sodium sulfate, yield high early strengths (e.g., 20-30 MPa in 24 hours) and are used in lightweight concrete masonry units or thermal insulation panels, where they replace up to 20% of Portland cement to reduce density without compromising load-bearing capacity. However, their sensitivity to moisture requires careful formulation to avoid delayed ettringite expansion, as patented compositions incorporate stabilizers like citric acid for dimensional stability.
PropertyAnhydrite ScreedTraditional Cement Screed
Drying Time to Walkable24 hours3-7 days
Thickness Range25-100 mm50-100 mm
Thermal Conductivity1.1-1.4 W/m·K1.2-2.0 W/m·K
Shrinkage<0.05%0.1-0.5%
This table compares key performance metrics, highlighting anhydrite's advantages in speed and finish quality for modern construction timelines.

Agricultural and Environmental Amendments

Anhydrite, or anhydrous (CaSO₄), functions as a soil amendment in agriculture by supplying calcium and sulfur, essential nutrients that enhance soil fertility and structure without significantly altering pH levels. Unlike (CaSO₄·2H₂O), anhydrite dissolves more slowly, providing a prolonged, timed-release effect that sustains nutrient availability to plant roots over extended periods. This gradual dissolution makes it particularly suitable for long-term soil conditioning in regions with evaporite deposits, where it can be sourced locally to reduce transportation costs. In sodic or alkaline soils, anhydrite aids reclamation by displacing sodium ions through calcium exchange, converting sodium carbonate to more soluble sodium sulfate and facilitating improved water infiltration and reduced surface crusting. It also promotes better soil aggregation, mitigating erosion and enhancing root penetration, which benefits crops like peanuts, alfalfa, and cotton that respond positively to sulfur supplementation—anhydrite contains approximately 25% more sulfur by weight than . Application rates typically range from 1 to 5 tons per hectare, depending on soil analysis, with efficacy demonstrated in field trials showing increased yields in sulfur-deficient soils. Environmentally, anhydrite's use as a soil conditioner supports sustainable land management by recycling industrial by-products, such as those from flue gas desulfurization, into amendments that minimize synthetic fertilizer dependency and reduce nutrient leaching risks due to its low solubility. It poses minimal ecological disruption, as it is non-toxic and biodegradable in soil systems, though slower nutrient release may limit its immediate effectiveness in acute remediation scenarios compared to more soluble alternatives. Limited studies indicate potential in stabilizing contaminated sites by binding heavy metals via sulfate complexation, but agricultural applications predominate over broader environmental remediation.

Significance and Implications

Paleoenvironmental Indicators

Anhydrite deposits primarily form through the precipitation of calcium sulfate from concentrated brines in evaporative settings, signaling paleoenvironments with elevated evaporation exceeding precipitation and restricted marine or lacustrine circulation. These conditions typically reflect arid to semi-arid climates, often associated with subtropical high-pressure belts, where seawater or hypersaline waters concentrate in isolated basins, sabkhas, or lagoons during marine regressions or tectonic restriction. Thick sequences of bedded or nodular anhydrite, interlayered with halite or carbonates, indicate prolonged hypersalinity and low clastic input, distinguishing them from gypsum-dominated deposits that favor cooler or less saline regimes. Fabrics and textures in ancient anhydrite provide diagnostic clues to specific depositional subenvironments and relative sea-level dynamics. Nodular, enterolithic, or chicken-wire structures often denote supratidal sabkha settings with periodic flooding and desiccation, while massive or laminated bedding suggests subtidal or deeper basinal precipitation under stratified brines. Replacement anhydrite, formed by dehydration of primary gypsum during burial or hot, saline diagenesis, further implies post-depositional warming or increased overburden pressure, common in rift or foreland basins. Such features, when correlated with cyclicity in evaporite cycles, reconstruct eustatic or autocyclic sea-level fluctuations, as seen in Miocene basins where anhydrite caps mark highstands in salinity. In broader paleoclimatic reconstructions, widespread anhydrite-bearing evaporites correlate with greenhouse intervals favoring aridity, such as the , where their distribution tracks paleolatitudinal belts of evaporation. Associated redox-sensitive elements and isotopic signatures (e.g., sulfur isotopes reflecting bacterial sulfate reduction) reveal basin oxygenation and brine density stratification, influencing organic matter preservation in adjacent shales. However, distinguishing primary marine from non-marine or hydrothermal origins requires integrating trace element geochemistry, as continental evaporites may mimic marine signals under extreme aridity. These indicators underscore anhydrite's role in delineating episodes of global or regional drying, tectonically driven isolation, and shifts in ocean-atmosphere hydrology.

Engineering Challenges and Geohazards

Anhydrite deposits pose significant engineering challenges due to their solubility in water and capacity for volumetric expansion during hydration to . Dissolution occurs rapidly under near-surface conditions or along discontinuities, forming cavities and karst-like features that can lead to subsidence and sinkhole collapses on timescales observable within human lifespans. In regions with , such as Ripon, North Yorkshire, England, natural gypsum dissolution—often involving associated anhydrite—has triggered catastrophic subsidence events approximately every three years, complicating infrastructure development and requiring extensive geotechnical mitigation. Hydration of anhydrite to gypsum, typically induced by exposure to groundwater or meteoric water, results in substantial expansion, with volume increases of up to 61% generating swelling pressures that uplift overlying strata and damage foundations, tunnels, and pipelines. This process exacerbates geotechnical instability in interbedded sequences with clays, where swelling propagates differentially, leading to differential settlement and structural distress in coastal or basin-margin deposits, as observed in Arabian Gulf regions. Geotechnical investigations reveal that anhydrite exhibits plastic-elastic-plastic deformation behavior under load, transitioning to earlier plastic flow upon partial hydration to gypsum, which reduces shear strength and increases compressibility in foundations. These properties necessitate specialized site assessments, including geophysical surveys and grouting, to predict and counteract hazards like cavity collapse or heave in evaporite karst terrains. In the United States, evaporite dissolution risks, including those from anhydrite, have prompted mapping of subsidence-prone areas to inform land-use planning and avoid unmitigated development over soluble bedrock.

Debates and Unresolved Questions

Gypsum-Anhydrite Precipitation Paradox

The gypsum-anhydrite precipitation paradox describes the difficulty in directly precipitating (CaSO₄) from aqueous solutions at low temperatures (<120 °C) in laboratory settings, despite its thermodynamic stability in certain conditions and prevalence in ancient evaporite deposits. In natural shallow marine or lacustrine evaporative environments, (CaSO₄·2H₂O) typically forms first upon CaSO₄ supersaturation, at evaporation factors of 3.3–3.8 times seawater, followed by diagenetic dehydration to anhydrite that preserves primary fabrics such as enterolithic or chickenwire structures. Thermodynamically, gypsum is the stable phase below the transition temperature of approximately 42–58 °C at 1 atm pressure, where solubilities converge at ~0.015 mol/kg H₂O; above this, anhydrite becomes stable with decreasing solubility. Density functional theory calculations confirm a transition at 46.58 °C, attributing gypsum's higher stability at ambient conditions to stronger ion-covalent bonding and a free energy of formation of -1849.72 kJ/mol versus -1350.79 kJ/mol for anhydrite. However, in hypersaline brines at surface temperatures (<40 °C), supersaturation thresholds favor gypsum precipitation, as its solubility (~0.015 mol/kg at 25 °C) allows earlier nucleation compared to anhydrite. Kinetically, anhydrite precipitation is hindered by high nucleation energy barriers and slow growth rates, requiring extended periods (>2 years at 60 °C) for formation even from supersaturated solutions or phase transitions from or bassanite (CaSO₄·0.5H₂O). Experimental modeling in NaCl solutions (0.8–4.3 M, 40–120 °C) demonstrates as the primary precipitate below 80 °C, with anhydrite emerging only secondarily over geological timescales, explaining its pseudo-primary occurrence in nodular or bedded evaporites without direct precursors. This kinetic constraint resolves the paradox, as direct low-temperature anhydrite formation remains rare, confined to warmer subsurface or long-residence-time settings where overcome barriers.

Dissolution and Stability Dynamics

Anhydrite (CaSO₄) is thermodynamically stable relative to (CaSO₄·2H₂O) above a of approximately 42°C ± 1°C in aqueous systems, as established through equilibria and calorimetric assessments. Below this threshold, becomes the stable phase, yet anhydrite's persistence in low- environments stems from kinetically hindered dehydration of and hydration of anhydrite, with direct conversion rates negligible below 90°C due to barriers. Phase diagrams of the CaSO₄-H₂O system illustrate anhydrite's curve intersecting 's at this , with anhydrite increasing positively with (e.g., from ~0.21 g/100 mL at 25°C to higher values near 100°C), enabling its dominance in subsurface deposits formed under elevated thermal gradients. further modulates stability, favoring anhydrite at ionic strengths exceeding ~3 m NaCl equivalents, where peaks before declining. Dissolution of anhydrite proceeds via a second-order law, contrasting gypsum's , with rates governed by surface reaction controls rather than in undersaturated solutions. Experimental data indicate dissolution rates accelerate with temperature, acidity (optimal near 4-5), and , but remain slower than gypsum due to anhydrite's denser and lower initial (e.g., ~2.1 × 10⁻³ / at 25°C versus gypsum's ~0.015 /). In brines, suppresses dissolution by common-ion effects, yet prolonged exposure in infiltration zones promotes partial retroconversion to via , altering pore structures and mechanical integrity. Geologically, these dynamics manifest in karstification, where anhydrite dissolution—though gradual—undermines integrity, inducing and pipe formation over millennia, as observed in basins like the Holbrook Basin, , with collapses spanning meters to kilometers. Human interventions, such as or tunneling, exacerbate instability by shifting local hydrogeochemistry toward undersaturation, precipitating infills that expand volumes by up to 60% and trigger delayed hazards. Such processes underscore anhydrite's role in long-term landscape evolution, with empirical monitoring in regions like the Ebro Valley revealing dissolution rates of 0.1-1 mm/year under natural recharge, amplifying risks in overlying aquifers and infrastructure.

References

  1. [1]
    Anhydrite Mineral Data - Mineralogy Database
    Help on Chemical Formula: Chemical Formula: CaSO4 ; Help on Composition: Composition: Molecular Weight = 136.14 gm ; Calcium 29.44 % Ca 41.19 % CaO ; Sulfur 23.55 ...
  2. [2]
    Anhydrite | Properties, Formation, Occurrence and Uses Area
    Apr 21, 2024 · Chemical Formula: CaSO4 – Anhydrite consists of calcium (Ca), sulfur (S), and oxygen (O) atoms. Water Content: Anhydrite is an anhydrous mineral ...
  3. [3]
    The mineral anhydrite information and pictures
    Anhydrite specimens in a collection may also alter to Gypsum if kept in moist conditions over a prolonged period of time. Chemical Formula. CaSO4. Color.Missing: polymorphs | Show results with:polymorphs
  4. [4]
    Calcium Sulfate | CaSO4 | CID 24497 - PubChem
    NATURAL FORM OF ANHYDROUS CALCIUM SULFATE IS KNOWN AS MINERAL ANHYDRITE; ALSO AS KARSTENITE, MURIACITE, ANHYDROUS SULFATE OF LIME, ANHYDROUS GYPSUM. ... Calcium ...
  5. [5]
    [PDF] Anhydrite CaSO4 - RRuff
    Occurrence: A major component in sedimentary evaporite deposits and in the cap rocks above salt domes, commonly formed by dehydration of gypsum; in igneous ...Missing: formula | Show results with:formula
  6. [6]
    ANHYDRITE - HPF Minerals
    The main properties of anhydrite​​ Anhydrite is known for its moisture resistance, medium hardness and chemical stability. It crystallizes in the orthorhombic ...Missing: occurrence | Show results with:occurrence<|separator|>
  7. [7]
    Anhydrite: Mineral information, data and localities.
    Isostructural with Ferruccite; isostructural and isomorphous with α-BaSO4 and α-SrSO4. Readily alters to gypsum.About Anhydrite · Chemistry · Chemical Analysis · Crystal Structure
  8. [8]
    [PDF] Some crystallographic studies in the system CaSOa-CaSO 4 ... - RRuff
    Summary. The existence and relationships of the polymorphs of CaSOa have been studied by X-ray methods. Only two distinct phases are found: no crystallo-.
  9. [9]
    [PDF] TEMPERATURE PHASE OF CALCIUM SULFATE IN ... - Ohio.gov
    The existence of a high-temperature polymorphic inversion of calcium sulfate (anhydrite) between 1190°C and 1210°C was suggested originally by Grahmann in.
  10. [10]
    The crystal structure of anhydrite (CaSO4) - IUCr Journals
    The structure of anhydrite (CaSO4) was determined by Wasastjerna in 1925 and a somewhat different solution for its structure was described by Dickson and Binks ...
  11. [11]
    mp-4406: CaSO4 (Orthorhombic, Cmcm, 63) - Materials Project
    CaSO₄ is Zircon-like structured and crystallizes in the orthorhombic Cmcm space group. Ca²⁺ is bonded in a 8-coordinate geometry to eight O²⁻ atoms.
  12. [12]
    New Insights into the Transition Temperature of Gypsum to ...
    Anhydrite belongs to the orthorhombic space group Cmcm. In the anhydrite structure, CaO8 dodecahedra and SO4 tetrahedra form alternating edge-sharing chains ...
  13. [13]
    Anhydrite Mineral | Uses and Properties - Geology.com
    Physical Properties of Anhydrite ; Streak, White ; Luster, Vitreous to pearly ; Mohs Hardness, 3 to 3.5.
  14. [14]
    10 Sulfates - Anhydrite - Optical Mineralogy
    Anhydrite has high birefringence; maximum interference colors are 3rd-order green. Polysynthetic twinning is commonly seen under crossed polars (Fig 10.1. 2). ...
  15. [15]
    KGS--Geological Log Analysis--Evaporites
    Mar 24, 2017 · As products of evaporation from sea-water, the minerals are deposited in a set order as was first demonstrated by Usiglio in 1884. He ...
  16. [16]
    [PDF] Lithology of Evaporite Cycles and - USGS Publications Warehouse
    Periodic, rapid precipitation would have caused a rain of gypsum (anhydrite) crystals onto bottom-growth crystals of halite. An example of this mechanism for ...
  17. [17]
    [PDF] THE GYPSUM-ANHYDRITE EQUILIBRIUM AT ONE ATMOSPHERE ...
    Unit cell parameters of the synthetic gypsum were determined by. X-ray powder diffraction methods with either silicon or quartz as internal standard, and ...
  18. [18]
    Evaporite rocks - ALEX STREKEISEN
    The gypsum and anhydrite grow by displacement within the sediment, with the gypsum in clusters and the anhydrite forming amorphous coalesced nodules with little ...
  19. [19]
    Mineralogy of evaporites: Marine basins - Geological Digressions
    Mar 19, 2020 · Evaporite precipitation mostly occurs at the brine surface where mineral supersaturation is maintained by evaporation. Floating crystal rafts ...
  20. [20]
    [PDF] the stability of gypsum and anhydrite in the geologic environment
    In the CaSO-- H2 0 system, gypsum and anhydrite are the only two minerals that are sufficiently stable to exist as geologic deposits.
  21. [21]
    An accurate model for the solubilities of anhydrite in H2O–NaCl ...
    Apr 5, 2024 · Anhydrite deposition can be triggered by decreased solubility and/or water evaporation. Temperature (T), pressure (P), and fluid composition ...
  22. [22]
    A genetic explanation for the anhydrite–halite cyclic layers in the ...
    Sep 6, 2024 · Halite precipitation in evaporite basins often begins as a near-surface phenomenon involving the formation of hopper crystals. They are common ...
  23. [23]
    [PDF] Evaporites through time - Saltwork Consultants Pty Ltd
    Nov 22, 2009 · Ongoing gypsum/anhydrite precipitation removes almost all of the Ca from a seawater brine leaving excess sulfate in the liquor. After halite ...
  24. [24]
    Gypsum-anhydrite transition: a source of water during diagenesis?
    It has been established [1] that the gypsum/anhydrite transition, accompanied by release of 480 kg of water per m3 of gypsum, occurs below 60ºC at a ...
  25. [25]
    [PDF] Petrography of Diagenetic Anhydrite in the Middle Devonian ...
    Anhydritization was one of the key diagenetic processes that influenced porosity distribution in the Winnipegosis and Ratner formations. Previous studies of ...
  26. [26]
    Anhydrite diagenesis in a vegetated sabkha, Al-Khiran, Kuwait ...
    In the Al-Khiran sabkha, supratidal anhydrite occurs in three distinct morphological forms: (a) hair-cream anhydrite which occurs exclusively in association ...
  27. [27]
    ANHYDRITE DIAGENESIS, CALCITIZATION, AND ORGANIC ...
    Mar 2, 2017 · ANHYDRITE DIAGENESIS ... Carbon isotope excursion and its genetic mechanism during the Sinian to Cambrian transition in the northern Tarim Basin.
  28. [28]
    Minerals Filling in Anhydrite Dissolution Pores and Their Origins in ...
    Apr 20, 2021 · In the present case, the authigenic quartz filling in anhydrite mold is, therefore, formed as a burial diagenetic process instead of an early ...
  29. [29]
    Emplacement and diagenesis of gypsum and anhydrite in the late ...
    Upon uplift of the Raisby Formation during the Tertiary, anhydrite rehydrated to porphyroblastic gypsum and this was then dissolved by meteoric groundwaters.
  30. [30]
    The isotopic record of gypsum diagenesis in diluted solutions
    Aug 20, 2018 · Both the increases of temperature, activity of water and pressure promote the conversion of gypsum to anhydrite, the anhydrous sulfate mineral ( ...Missing: dehydration | Show results with:dehydration
  31. [31]
    Tracing the chemical evolution of fluids during hydrothermal recharge
    Anhydrite becomes less soluble in seawater with increasing temperature, and precipitates from seawater at temperatures >150°C at 500 bar pressure [10].
  32. [32]
    Genesis of anhydrite in hydrothermally altered basalt from the East ...
    Jan 7, 2013 · The existence of anhydrite in the altered basalt indicates extensive high-temperature hydrothermal alteration at the surface of seafloor pillow ...
  33. [33]
    Formation of a hydrothermal reservoir due to anhydrite precipitation ...
    Nov 16, 2010 · Anhydrite precipitation strongly affects the circulation of hydrothermal systems by changing the permeability structure of the oceanic crust ...
  34. [34]
    [PDF] The control of anhydrite sulfidation on the stockwork mineralization ...
    Our results suggest that - at stockwork depth - anhydrite is removed by the interaction with anhydrite-undersaturated hydrothermal fluids, leaving behind pores ...
  35. [35]
    Anhydrite‐Assisted Hydrothermal Metal Transport to the Ocean ...
    Jun 2, 2020 · The establishment of such anhydrite-sealed zones reduces mixing between the hydrothermal fluid and seawater and results in an increase in vent ...
  36. [36]
    Microbially influenced formation of anhydrite at low temperature
    Dec 1, 2023 · We present the results of laboratory precipitation experiments showing that anhydrite can form at 35 °C from evaporated seawater through a microbially ...
  37. [37]
    (PDF) Microbially influenced formation of anhydrite at low temperature
    Aug 7, 2025 · Here, we present the results of laboratory precipitation experiments showing that anhydrite can form at 35°C from evaporated seawater through a ...
  38. [38]
    Desert Microbes Mine for Water - Eos.org
    Jun 29, 2020 · Once the water was extracted, the mineral components left behind precipitated into new nanosized crystals of anhydrite.
  39. [39]
    Mechanism of water extraction from gypsum rock by desert ...
    May 4, 2020 · Upon losing the water of crystallization, the monoclinic gypsum crystals become unstable and transform to orthorhombic anhydrite crystals. The ...
  40. [40]
    Effect of bacteria and salinity on calcium sulfate precipitation
    Feb 15, 2012 · This paper presents interactive effects of biological matter, Pseudomonas Fluorescence (PF), on precipitation of calcium sulfate, a sparingly soluble salt and ...
  41. [41]
    [PDF] LOW TEMPERATURE ANHYDRITE PRECIPITATION IN FLOWING ...
    Instead, anhydrite may re- quire a constant source of low-water activity brine and/or advective solute transport to form at low tem- peratures. Dissolution ...
  42. [42]
    The gypsum–anhydrite paradox revisited - ScienceDirect.com
    Oct 29, 2014 · ... minimum temperature for direct anhydrite precipitation should be even further lowered, to about 18 °C. However, the stability region of each ...
  43. [43]
    7.1: Evaporites - Geosciences LibreTexts
    Mar 9, 2025 · The two are closely associated and with increased burial depth gypsum commonly loses the water and transforms into anhydrite. The process can ...
  44. [44]
    Tectonic Controls on Evaporite - SEPM Strata
    Mar 11, 2015 · Thick sections of evaporites (anhydrite, gypsum and halite) in the geologic record usually have accumulated adjacent to margins of recently ...
  45. [45]
    KGS--Bulletin 169--Haney and Briggs - Kansas Geological Survey
    Abstract. Evaporites are characterized in a vertical sequence by cycles of rock types, such as dolomite-anhydritehalite-anhydrite-dolomite. Associated with the ...
  46. [46]
    Basinal Evaporite Depositional Environments
    The Permian Castile Formation in West Texas is an excellent example of basinal evaporite produced in a barred basin. Laminated Permian Castile evaporite from ...
  47. [47]
    Sedimentological and diagenetic patterns of anhydrite deposits in ...
    Anhydrite deposits are widely distributed in the Middle Miocene Badenian evaporite basin of Poland, including the marginal sulphate platform and adjacent salt ...
  48. [48]
    Giant polygonal anhydrite ridges in the Southern Permian Basin
    Oct 30, 2024 · Giant polygonal anhydrite ridges with diameters of 2–4 km have been identified in the NE margin of the Southern Permian Basin in Poland.<|separator|>
  49. [49]
    Carbonate formation in salt dome cap rocks by microbial anaerobic ...
    Feb 18, 2019 · Approximately 65% of the onshore salt domes in the GMB are mantled by thick (up to 300 m) cap rock. This cap rock consists of anhydrite, gypsum, ...<|separator|>
  50. [50]
    Cyclicity of katatectic layers in anhydrite caprocks, U.S. Gulf Coastal ...
    Jul 15, 2024 · In the anhydrite caprock of Winnfield Salt Dome, a layer of sulfides (dominantly pyrrhotite) uncommonly formed a laminar zone parallel to the ...
  51. [51]
    Multiple fluid components of salt diapirs and salt dome cap rocks ...
    Salt diapirs contain a few percent of anhydrite that accumulated as residue to form anhydrite cap rocks during salt dissolutions. Reported 87Sr/86Sr ratios ...
  52. [52]
    Textural Evidence for Origin of Salt Dome Anhydrite Cap Rocks ...
    Textures within anhydrite cap rock are products of repeated cycles of halite dissolution and residual anhydrite accretion at tops of salt stocks.
  53. [53]
    [PDF] ORIGIN OF THE ANHYDRITE GAP ROCK OF AMERICAN SALT ...
    ORIGIN OP ANHYDRITE CAP ROCK OF AMERICAN SALT DOMES. 85 vening more plastic salt 9 ; or it may have accumulated within the stock by solution of the salt.
  54. [54]
    [PDF] Holistic Understanding of the Big Hill Salt Dome Geologic Processes ...
    Rock type: comprised of a lower anhydrite zone overlain by a limestone and gypsum zone. Top of Cap: depth of ~ 300 feet. Thickness: between 850-1300 feet !
  55. [55]
    [PDF] Chapter 12 Salt Domes in the Gulf Coast Aquifer
    Cap rock formation results from salt dissolution. Anhydrite. (calcium sulfate), the main impurity in the salt stock, forms a residual accumulation at the dome.
  56. [56]
    [PDF] Mississippi Geology
    Caprock at Tatum Dome exhibits characteristics typical of zoned caprocks from other Gulf Coast salt domes (Halbouty,. 1979). The caprock is 500-600 feet thick ...
  57. [57]
    [PDF] Descriptive Model of Salt-Dome Sulfur
    Salt-dome sulfur deposits are biogenic sulfur deposits that form in the anhydrite-gypsum caprocks of salt domes. They differ from other.<|control11|><|separator|>
  58. [58]
    The Structure and Significance of Anhydrite-Bearing Vein Arrays ...
    Aug 7, 2025 · A spectacular anhydrite ± carbonate vein array is exposed in the deeper levels of the Lienetz open pit and reveals a dynamic structural ...
  59. [59]
    [PDF] ANHYDRITE AND GYPSUM IN THE LYON MOUN
    The occurrence of the minerals anhydrite and gypsum at the Chateaugay Mine at. Lyon Mountain, New York is described. The anhydrite is believed to be ...Missing: notable | Show results with:notable
  60. [60]
    (PDF) Origin and evolution of a salty gypsum/anhydrite karst spring
    It drains an aquifer of unique properties composed of anhydrite with halite lenses at depth and gypsum at the surface (both with high NaCl content).Missing: hydrothermal | Show results with:hydrothermal
  61. [61]
    [PDF] ANHYDRITE var - Celestial Earth Minerals
    Anhydrite also occurs to a lesser extent in carbonatites, zeolite pockets within basalt formations, hydrothermal veins, altered limestone, and volcanic ...
  62. [62]
    Anhydrite - the Dry Gypsum - MineralExpert.org
    Dec 9, 2018 · Physical Properties of Anhydrite​​ Anhydrite is colorless to pale blue. It tends to violet if transparent; impurities such as Fe or Mn will tint ...Missing: formula | Show results with:formula
  63. [63]
    Anhydrite - ALEX STREKEISEN
    Anhydrite is a mineral anhydrous calcium sulfate, CaSO4, named for its lack of water of crystallization, and it transforms to gypsum when exposed to water.
  64. [64]
    Anhydrite - Gemstone Dictionary
    Origin of name: firstly discovered in 1794 by Austrian entomologist and mineral collector Nicolaus Poda von Neuhaus, who called the mineral muriacite ...
  65. [65]
    Genesis of primary structures in anhydrite - GeoScienceWorld
    Mar 3, 2017 · Anhydrite, a common evaporite in all systems since the Cambrian, displays a variety of structures which may be classified as layered, ...
  66. [66]
    ANCIENT ANHYDRITE FACIES AND ENVIRONMENTS, MIDDLE ...
    supratidal, shallow-water, deep-water, and replacement. Supratidal ...
  67. [67]
    (PDF) The problem of distinguishing between primary and ...
    Nov 30, 2015 · PDF | On Jan 1, 1985, L.A. Hardie and others published The problem of distinguishing between primary and secondary features in evaporites ...Missing: key | Show results with:key
  68. [68]
    Exploitation that focuses on excellence - Knauf - L'anhydrite Lorraine
    The underground working of the quarry follows the method of abandoned pillars and chambers. After blasting by drilling and explosive firing, the loader- ...<|separator|>
  69. [69]
    Anhydrite gypsum | Danucem – built on responsibility
    Raw material extraction consists of the following activities: drilling; blasting; removal; crushing; storage shipment. Shipping method. We offer natural ...
  70. [70]
    [PDF] Gypsum and Anhydrite - NERC Open Research Archive
    Jul 23, 2025 · Commercial deposits are confined to the Upper Permian of Cumbria and North Yorkshire, the Keuper Marl of Nottingham- shire and Staffordshire ...
  71. [71]
    Anhydritec - Minersa Group - Anhydrite
    Although anhydrite can be found in natural mineral deposits, it can also be obtained as a byproduct from different industrial processes. The Minersa Group ...Missing: extraction | Show results with:extraction
  72. [72]
    Anhydrite Market Growth Outlook (2023 to 2033) - Fact.MR
    The United States is also a major producer of anhydrite, with large deposits located in states such as New York, Michigan, and New Mexico. In December 2020 ...
  73. [73]
    [PDF] Occurrence and Stratigraphy of some Gypsum and Anhydrite ...
    Location of gypsum and anhydrite deposits in Alberta. 9. Location of the Mowitch Creek and the Fetherstonhaugh. Creek Triassic gypsum deposits. -opposite 10.<|separator|>
  74. [74]
    [PDF] Report (pdf) - USGS Publications Warehouse
    Tertiary carbonate-anhydrite sequences into the South Florida basin 29. 2.17. Generalized geologic column of Lower Cretaceous strata in the South. Florida basin ...
  75. [75]
    Anhydrite Market - Size, Share, Industry Trends, and Forecasts (2025
    Rating 5.0 (186) May 20, 2025 · Additionally, the easy extraction and processing of natural anhydrite reduce manufacturing costs, making it a preferred choice for large-scale ...
  76. [76]
    Influence of Anhydrite on the Properties and Microstructure of ... - NIH
    Oct 9, 2022 · The results show that adding an appropriate amount of anhydrite promotes the hydration of APC. The highest compressive strength is reached at an anhydrite ...
  77. [77]
    The Anhydrite Process for Sulfuric Acid and Cement
    In this process, anhydrite (calcium sulfate) replaces limestone in a cement rawmix, and under reducing conditions, sulfur dioxide is evolved instead of carbon ...
  78. [78]
    Production of Anhydrite Binder from Waste Fluorangydrite - MDPI
    Mar 27, 2023 · The process of recycling fluorine hydrite waste from the production of hydrofluoric acid at Ulba Metallurgical Plant JSC was studied, and anhydrite unburnt ...<|separator|>
  79. [79]
    Making of anhydrite cement from waste gypsum - ScienceDirect.com
    The paper presents the results of research on the utilization of phosphogypsum produced as the waste of phosphoric acid manufacture.
  80. [80]
    Optimization of the activation process for natural anhydrite
    Chemical activation of anhydrite is set out exclusively by dosage of activators of different nature: ground granulated blast furnace slag, ground glass residue, ...
  81. [81]
    Anhydrite CaSO4 supplier - DDFluor
    In construction, anhydrite is used as a binder for the production of self-levelling screeds, replacing cement. This has many advantages, such as better fluidity ...
  82. [82]
    Mastering Anhydrite Screed: A Guide to Successful Tiling - RUBI
    Jun 21, 2024 · Anhydrite screed, also known as calcium sulphate screed, is a type of self-levelling compound used to create a smooth surface for flooring.
  83. [83]
    Anhydrite Screed vs. Cement: Which Is the Better Choice?
    Dec 13, 2024 · Anhydrite dries quickly, good for underfloor heating, but needs moisture monitoring. Cement is durable, moisture-resistant, but slower to cure, ...Drying Time Differences · Cement Drying Variability · Moisture Sensitivity
  84. [84]
    Installing floors over anhydrite, calcium sulphate and gypsum screeds
    Anhydrite screeds, also referred to as calcium sulphate or gypsum screeds are becoming more prevalent because they provide a cost effective and eco-friendly ...
  85. [85]
    Preparing Anhydrite Screed for Wood Flooring
    Anhydrite screeds – also referred to as calcium sulphate or gypsum-based screeds – are a cost-effective, eco-friendly alternative to traditional sand and cement ...<|separator|>
  86. [86]
    The Influence of Anhydrite on the Mechanical Performance of ...
    Mar 29, 2025 · This study investigates the sample of the grouting material with (calcium sulfoaluminate cement, anhydrite, and quicklime) under different ratios.
  87. [87]
    New insights on the action mechanisms of anhydrite on autogenous ...
    Aug 8, 2025 · Therefore, anhydrite (AH) was selected as an additive for preparing the CL-UHPC matrix. The fresh performances (mini slump flow and setting time) ...
  88. [88]
    Making of anhydrite cement from waste gypsum - ScienceDirect
    1. The stable anhydrite cement can be produced by heating phosphogypsum at 1000°C. 2. The microstructure of anhydrite cement shows the formation of euhedral ...Missing: production | Show results with:production
  89. [89]
    [PDF] Low energy synthesis of anhydrite cement from waste lime mud
    Dec 8, 2022 · 4-8 The traditional technology for pro- ducing anhydrite cement consists of firing gypsum stone in rotary kilns at a temperature of 800-1000 °С, ...
  90. [90]
    [PDF] The influence of anhydrite III as cement replacement material in ...
    Dec 4, 2022 · Lightweight concrete masonry units are mainly used to construct walls as a building material. To produce nor- mal-weight concrete masonry units, ...
  91. [91]
    Structural and Thermal Insulation Materials Based on High-Strength ...
    The use of anhydrite compounds in the manufacture of products is limited with low speed setting and hardening. To activate the structure formation of the ...
  92. [92]
    Anhydrite cement compositions and procedures for manufacture
    The present invention describes improved anhydrite cementitious compositions that control or avoid ettringite formation, as well as procedures for manufacture.
  93. [93]
    A Comprehensive Guide to Anhydrite Floor Screed
    Apr 2, 2024 · Anhydrite screeds can be laid more quickly than traditional screeds, reducing construction time. It can be pumped over large areas quickly, ...
  94. [94]
    Anhydrite - an overview | ScienceDirect Topics
    Minerals with the same chemical formula or chemical composition but different structures are called polymorphs, e.g., calcite (CaCO3) and aragonite (CaCO3); in ...
  95. [95]
    Calcium Sulfate - 09 Which Gypsum: Anhydrite or Dihydrate?
    Nov 23, 2020 · It can release a fast burst of calcium and sulfur into the soil to restore balance and bring up the base levels of nutrients in the soil.
  96. [96]
    The use of anhydrite as a fertiliser
    As anhydrite contains 25% more sulphur than gypsum, it is economically advantageous to convert gypsum into anhydrite before transport over considerable.
  97. [97]
    Arcosa Specialty MaterialsGypsum Anhydrite and Gypsum Dihydrate
    It is also a soil treatment, useful both for modifying soil chemistry and improving soil structure. Calcium also decreases surface crusting and the surface ...
  98. [98]
    [PDF] USING GyPSUM AND OTHER CALCIUM AMENDMENTS
    In all soils, regardless of pH, gypsum is a good Ca2+ additive when Ca2+ is needed. Calcium sulfate anhydrite will also supply Ca2+, but it will dissolve more ...
  99. [99]
    Benefits to Specific Crops: Gypsum, Anhydrite, & Limestone
    Limestone, anhydrite and gypsum have been used for centuries as an additive to enrich the soil with missing minerals, correct pH values and provide other ...
  100. [100]
    The Many Benefits of Gypsum, Anhydrite, and Limestone
    Increased Water Infiltration · Crust Prevention · Better Moisture Retention ; Fertilization · Enhanced Nutrient Absorption · Environmentally Safe ; Irrigation Effect ...Missing: environmental | Show results with:environmental
  101. [101]
    A fine diversity of applications - Knauf - L'anhydrite Lorraine
    Natural anhydrite is used as a soil modifier, and is well known to cement manufacturers, who use it to slow down setting.
  102. [102]
    Evaporites and red beds of a syn-rift Atlantic series - GeoScienceWorld
    Nov 6, 2024 · Evaporites can be used to interpret paleoenvironments and to reconstruct depositional settings. In particular, ancient evaporites are indicators ...
  103. [103]
    17 The Jurassic Climate | Climate in Earth History
    CLIMATIC CRITERIA​​ Substantial deposits of evaporites (notably gypsum, anhydrite, and halite) indicate conditions of both warmth and aridity, whereas coals ...
  104. [104]
  105. [105]
    Anhydrite fabrics as an indicator for relative sea-level signatures in ...
    The evaporite-bearing intervals provide an opportunity to investigate the interactions between sea-level fluctuations and the growth of anhydrites during the ...
  106. [106]
    Paleoenvironment of the Lower–Middle Cambrian Evaporite Series ...
    Jun 22, 2021 · The evaporites are mostly composed of halite (Figure 5a), anhydrite (Figure 5b), and mixed sediments with at least 50% anhydrite (Figure 5c,d).
  107. [107]
    On the Significance of Evaporites - NASA ADS
    ... anhydrite-halite-potash evaporites were formed by evaporation of ancient seawater. ... precipitation of anhydrite and calcite, and reduction of sulfate to H2S.<|separator|>
  108. [108]
    Gypsum/anhydrite: some engineering problems
    Gypsum and anhydrite may experience dissolution in the near-surface zone and along the discontinuities, creating karst terrain to depths related to present and/ ...Missing: challenges geohazards
  109. [109]
    Chapter 16 Geohazards caused by gypsum and anhydrite in the UK
    Jun 9, 2020 · The hydration of anhydrite to gypsum in the subsurface causes expansion and heave that are problematic to engineering and hydrogeological ...
  110. [110]
    Subsidence hazards caused by the dissolution of Permian gypsum ...
    About every three years natural catastrophic subsidence, caused by gypsum dissolution, occurs in the vicinity of Ripon, North Yorkshire, England.Missing: risks | Show results with:risks
  111. [111]
    Gypsum geohazards and sinkholes - NERC Open Research Archive
    Oct 20, 2021 · The natural or induced hydration of anhydrite to gypsum is accompanied by considerable expansion that can cause engineering problems for schemes ...Missing: challenges | Show results with:challenges
  112. [112]
    Chapter 16 Geohazards caused by gypsum and anhydrite in the UK
    In addition, the hydration and recrystallisation of anhydrite to gypsum can cause considerable expansion and pressures capable of causing uplift and heave.
  113. [113]
    Geotechnical Characteristics of Anhydrite/Gypsum Transformation in ...
    Nov 3, 2013 · The cracks in the inter-bedded clays are filled with evaporite materials indicating arid climatic conditions during exposure episodes. Swelling ...Missing: indicators | Show results with:indicators
  114. [114]
    Study on the geological and engineering aspects of anhydrite ...
    Swelling and compressibility problems are common in several coastal deposits of the globe. This paper investigates the geological and engineering aspects of ...
  115. [115]
    A survey of the engineering properties of some anhydrite and ...
    Most anhydrite and gypsum exhibit plastic-elastic-plastic deformation, subsequent plastic deformation occurring at an earlier stage during loading of gypsum ...Missing: challenges | Show results with:challenges<|separator|>
  116. [116]
    Subsidence hazards due to evaporite dissolution in the United States
    May 18, 2005 · If the dissolution cavity is large enough and shallow enough, successive roof failures above the cavity can cause land subsidence or ...Missing: risks | Show results with:risks
  117. [117]
    The gypsum–anhydrite paradox revisited - ScienceDirect.com
    Oct 29, 2014 · We show that intrinsic thermodynamic and kinetic properties severely constrain the precipitation of anhydrite (compared to gypsum and bassanite) ...Missing: precipitation | Show results with:precipitation
  118. [118]
    Gypsum Precipitation under Saline Conditions: Thermodynamics ...
    Gypsum is one of three possible calcium sulfate minerals and solutions from which it precipitates are typically oversaturated with additional minerals ...
  119. [119]
    Solubility of anhydrite and gypsum at temperatures below 100°C ...
    Sep 20, 2023 · The answer, how anhydrite could be formed at T < 50°C or 60°C is still open, since in the lab in time scales of years no primary precipitation ...
  120. [120]
    Temperature Dependence of Mineral Solubility in Water. Part 3 ...
    May 30, 2018 · Figure 5 shows a comprehensive overview of the solubility of the three relevant phases of calcium sulfate minerals (gypsum, hemihydrate, and ...
  121. [121]
    (PDF) Dissolution and conversions of gypsum and anhydrite
    Aug 5, 2015 · The dissolution of gypsum follows a first-order rate law, whereas the dissolution rate of anhydrite obeys a second-order law. Rates of gypsum ...
  122. [122]
    Mechanism and kinetics of gypsum–anhydrite transformation in ...
    Below 100°C, gypsum transforms directly to anhydrite. Above 100°C, it forms hemihydrate first. Increasing temperature and acid accelerates the process.Missing: transition | Show results with:transition
  123. [123]
    Reaction pathways and textural aspects of the replacement of ...
    Jun 1, 2017 · Since the solution layer closer to the calcium sulfate mineral surface becomes rapidly supersaturated with respect to any CaCO3 polymorphs, CaCO ...<|control11|><|separator|>
  124. [124]
    [PDF] National Evaporite Karst--Some Western Examples
    The Holbrook Basin in east-central Arizona demonstrates that dissolution of deeply buried evaporites can cause subsidence of overlying non- soluble rocks.
  125. [125]
    On the Solubilities of Anhydrite and Gypsum - SpringerLink
    During tunnel excavation, a change in the field conditions occurs, causing the dissolution of anhydrite into its ions and the subsequent precipitation of gypsum ...
  126. [126]
    Evaluation of geochemical and hydrogeological processes by ...
    The Ebro Valley in the outskirts of Zaragoza (NE Spain) is severely affected by evaporite karstification, leading to multiple problems related to subsidence ...Geological Setting And... · Selection Of Water Points... · Acknowledgements