Fact-checked by Grok 2 weeks ago

Subsidence

Subsidence is the gradual settling or sudden sinking of the Earth's surface owing to subsurface movement of earth materials. This phenomenon arises primarily from the removal or rearrangement of underground materials, leading to compaction or void formation. While natural processes such as tectonic shifts, glacial isostatic adjustment, and dissolution of soluble bedrock like limestone contribute, human activities—particularly excessive groundwater pumping, underground mining, and hydrocarbon extraction—dominate in many regions, causing irreversible aquifer compaction. Subsidence manifests at rates from millimeters to meters per year, inflicting substantial damage to infrastructure including buildings, roads, bridges, and utilities, while exacerbating flood risks by lowering land relative to sea level. Empirical measurements from interferometric synthetic aperture radar and ground-based monitoring reveal widespread occurrence, with over 20% of urban areas in major U.S. cities affected, impacting millions through fissuring, structural failures, and heightened vulnerability to inundation. Notable historical cases include subsidence exceeding 9 meters in California's San Joaquin Valley due to prolonged groundwater withdrawal, and episodic collapses from mining voids, underscoring the causal primacy of fluid extraction over gradual natural settling in populated zones.

Definition and Fundamentals

Geological Definition

Subsidence in geology denotes the gradual settling or abrupt sinking of the Earth's surface due to subsurface movement or displacement of earth materials, often involving compaction, dewatering, or removal of underlying strata. This process alters the elevation of land relative to surrounding areas and can affect areas from localized depressions to broad regional basins, with vertical displacements ranging from millimeters to several meters annually depending on the underlying geology. Geologically, subsidence differs from superficial soil settlement under immediate surface loads by originating from deeper lithospheric responses, such as the inelastic compaction of porous sediments or the dissolution of evaporites and carbonates, which reduce support for overlying layers. Natural examples include tectonic subsidence in sedimentary basins, where crustal loading leads to flexural downwarping, as observed in the Gulf of Mexico basin with rates up to 1-2 mm/year prior to significant human influence. Such movements reflect the dynamic interplay of rock mechanics and fluid dynamics within the Earth's crust, independent of anthropogenic factors.

Physical Mechanisms

The primary physical mechanism underlying subsidence is the compaction of subsurface sediments and soils, wherein pore spaces between particles contract under elevated effective stress, reducing overall volume and causing overlying materials to descend. This process predominantly affects compressible, fine-grained deposits such as clays and silts, where initial porosity can exceed 50-70%, allowing substantial deformation before reaching a stable density. Compaction may occur elastically (reversible) under low stress but transitions to inelastic behavior, yielding permanent subsidence as particles rearrange and interlock. In saturated porous media, consolidation drives this compaction according to Terzaghi's effective stress principle, which posits that the stress transmitted through the soil skeleton (effective stress, σ') equals total overburden stress (σ) minus pore fluid pressure (u): σ' = σ - u. Fluid withdrawal or loading decreases u, thereby increasing σ' and inducing shear and volumetric strains; excess pore pressures dissipate via drainage, expelling water and compressing the matrix over time scales from days to decades, depending on hydraulic conductivity. Low-permeability layers (aquitards) exhibit lagged, irrecoverable consolidation due to their high compressibility, with virgin compression indices often ranging from 0.2 to 1.0 in Holocene sediments, amplifying subsidence magnitudes up to meters in extreme cases. Primary consolidation dominates initial volume loss, followed by secondary compression from viscous particle sliding under sustained load. Subsidence can also arise from void formation without primary compaction, such as through subsurface dissolution of soluble minerals (e.g., carbonates, evaporites), eroding material via piping, or plastic flow in ductile strata like salt, culminating in gravitational collapse of roof spans when support thresholds are breached. Hydrocompaction represents a distinct rapid mechanism in unsaturated, low-density soils (e.g., loess with void ratios >1.0), where wetting eliminates metastable cementation, triggering sudden 1-5 m settlements as intergranular bonds fail. These processes collectively reflect causal linkages between stress perturbations, material properties, and kinematic failure, with surface expression varying by depth and heterogeneity.

Natural Causes

Tectonic and Seismic Processes

Tectonic subsidence arises from the deformation of the Earth's lithosphere driven by plate movements, primarily manifesting in sedimentary basins where the crust accommodates stresses through thinning, flexure, or thermal contraction. Key mechanisms include lithospheric extension, which reduces crustal thickness and induces isostatic adjustment, leading to rapid initial subsidence rates often exceeding 1 mm/year in active rifts; post-rift thermal cooling, where conductive heat loss causes contraction and slower subsidence over 10-100 million years; and loading by sediments or overriding thrust sheets in foreland basins, which flexes the lithosphere downward via isostatic compensation. These processes dominate in divergent margins, such as passive continental margins where initial syn-rift subsidence transitions to thermally driven subsidence, with total basin depths reaching kilometers over geological timescales. In convergent settings, subsidence occurs through flexural downwarping ahead of orogenic belts, as lithospheric loading from accreted sediments or thrust wedges depresses foreland basins; for example, subsidence rates in such basins can average 0.1-0.5 mm/year during active compression, accumulating to hundreds of meters over tens of millions of years. Complex interactions, such as in intracontinental basins, may combine these mechanisms without simple flexural models fitting observed patterns, as evidenced by variable subsidence across forelands influenced by inherited crustal weaknesses. Unlike anthropogenic subsidence, tectonic variants operate independently of human extraction, though they can amplify risks in populated basins when compounded with other factors. Seismic processes induce sudden subsidence via co-seismic slip on faults, where rupture displaces hanging walls downward relative to footwalls, often in extensional or thrust regimes. In subduction zones, megathrust earthquakes can cause broad coastal subsidence through underthrusting of the subducting plate, with vertical offsets of 1-3 meters documented in events like the Mw 9.2 1964 Alaska earthquake, where slip lowered sections of the overriding plate, exacerbating inundation in areas such as Cook Inlet. Similarly, the 1987 Mw 6.3 Edgecumbe earthquake in New Zealand produced localized subsidence of up to several decimeters via fault movement at depths around 8 km, demonstrating how seismic energy release triggers immediate tectonic adjustment. These events contrast with gradual tectonic subsidence by their abrupt nature, often followed by viscoelastic rebound, and primarily affect fault-proximal zones though far-field effects can occur via stress perturbations. In non-subduction contexts, such as strike-slip or normal faulting, subsidence manifests as graben formation or half-graben tilting, with magnitudes scaling to earthquake size and fault geometry.

Isostatic Adjustments

Isostatic adjustments refer to the vertical movements of the Earth's lithosphere in response to changes in surface or subsurface mass distribution, seeking to restore gravitational equilibrium as described by the principle of isostasy. In the context of subsidence, these adjustments occur when added loads, such as sediment accumulation or water bodies, increase the weight on the crust, causing it to sink until balanced by displacement of underlying mantle material. This process amplifies subsidence in depositional environments, where the infilling material itself contributes to further depression of the basin floor. Sediment loading exemplifies a primary natural mechanism of isostatic subsidence, particularly in foreland basins and deltas where erosion from uplands supplies thick accumulations that load the lithosphere. For instance, in tectonic settings like fold-thrust belts, the deposition of sediments can lead to flexural subsidence modulated by isostatic compensation, with rates influenced by sediment density and basin geometry; in ancient examples such as the Pliocene-Quaternary sequences on passive margins, seaward tilting subsidence has reached 240 meters per million years near shelf breaks due to combined loading and readjustment. Compaction of porous sediments further enhances this effect, reducing pore space and increasing effective density, thereby prompting additional sinking to maintain equilibrium. Thermal subsidence represents another isostatic process, driven by cooling and densification of the lithosphere following rifting or extension, which alters density contrasts and causes gradual sinking without external loading. This is evident in passive continental margins, where post-rift phases exhibit subsidence rates tied to conductive heat loss, often accumulating thousands of meters of sediment over tens of millions of years as the crust adjusts downward. Glacial isostatic adjustment (GIA), stemming from the Pleistocene ice sheet retreat approximately 15,000–20,000 years ago, primarily induces uplift in deglaciated cores but subsidence in peripheral regions through collapse of the peripheral forebulge—a raised rim formed by mantle flow away from the ice load. Along the U.S. East Coast, this manifests as ongoing subsidence, with rates up to 1.5 millimeters per year near New Jersey and contributing 1–3 millimeters per year in areas like the southern Chesapeake Bay region, exacerbating relative sea-level rise. In Virginia, GIA influences land motion persisting since the Laurentide Ice Sheet's maximum extent, though the region was unglaciated, resulting in subsidence rates of 1.1–4.8 millimeters per year observed since the 1940s. Similar patterns occur along Europe's Atlantic coast, where GIA models quantify contributions to subsidence on the order of millimeters per year.

Dissolution and Karst Formation

Dissolution of soluble bedrock, primarily limestone and dolomite composed of calcite (CaCO₃), occurs when groundwater slightly acidified by dissolved carbon dioxide (forming carbonic acid, H₂CO₃) reacts chemically to produce soluble calcium bicarbonate (Ca(HCO₃)₂), enlarging fractures and creating subsurface voids over geologic timescales. This process is most pronounced in carbonate rock formations where water flow is concentrated along joints and bedding planes, with dissolution rates typically ranging from 0.01 to 1 millimeter per year depending on water chemistry, flow velocity, and rock solubility, though cumulative effects span thousands to millions of years. Gypsum and evaporites like halite dissolve more rapidly due to higher solubility, potentially forming voids in decades to centuries under natural conditions. Karst landscapes emerge as dissolution propagates, yielding distinctive surface features such as closed depressions (dolines or sinkholes), blind valleys, and underground drainage systems, with aquifers exhibiting high permeability and rapid flow rates up to several meters per second in conduits. In bare karst (exposed bedrock), solutional sinkholes form directly through surface etching, while covered karst (overlain by soil or insoluble layers) develops suffosion sinkholes via soil piping into underlying voids or cover-collapse sinkholes from sudden roof failure. Subsidence manifests as gradual sagging of overburden into enlarged cavities or catastrophic collapse when structural integrity is lost, often triggered by natural variations in water table levels or increased infiltration during heavy rainfall, leading to surface displacements of meters to tens of meters in diameter and depth. Mechanisms include soil suffosion (internal erosion), sagging (plastic deformation), and brittle collapse, operating individually or combined based on cover thickness and bedrock type. Notable natural examples include the Edwards Plateau in Texas, where dissolution of Cretaceous limestone has produced over 30,000 documented sinkholes, some exceeding 100 meters in diameter, contributing to ongoing subsidence in karst-dominated watersheds. In the UK, Chalk and Jurassic limestones underlie regions like the Yorkshire Dales, where dissolution has formed subsidence sinkholes up to 20 meters deep, with historical collapses documented since the 18th century. Evaporite karst in the Permian Basin of west Texas illustrates faster subsidence, with salt dissolution creating depressions that enlarge at rates sufficient for surface features to develop within centuries, as evidenced by paleokarst remnants. These processes underscore karst subsidence as a primarily endogenous hazard, driven by long-term geochemical equilibrium shifts rather than external loading, though monitoring remains challenging due to subsurface concealment.

Anthropogenic Causes

Groundwater Extraction

Excessive extraction of groundwater from aquifers leads to land subsidence primarily through the compaction of unconsolidated sediments, such as clays and silts, within aquifer systems. When pumping reduces pore water pressure, the effective stress on the sediment grains increases, causing grains to rearrange and expel water, resulting in consolidation. This process is often irreversible for the initial "virgin" compression phase, where sediments undergo their first significant loading, leading to permanent loss of aquifer storage capacity. In confined aquifers overlain by aquitards, drawdown propagates upward, compressing fine-grained layers that contribute most to subsidence due to their high compressibility. Coarse-grained sands and gravels compact less because drainage allows quicker pore pressure dissipation. Subsidence manifests as differential sinking over pumping cones, exacerbating infrastructure damage and increasing flood vulnerability in coastal areas through relative sea-level rise. Notable examples include California's San Joaquin Valley, where groundwater pumping from the 1920s to 1970s caused subsidence exceeding 9 meters (30 feet) in some locations, damaging canals and aquifers. Recent InSAR monitoring reveals ongoing subsidence rates up to several centimeters per year in parts of the valley, driven by drought-induced pumping surges. In Italy's Emilia-Romagna region, historical over-pumping since the 1950s resulted in subsidence rates of 2-10 cm/year, linked to industrial and agricultural withdrawals. Globally, such extraction contributes to subsidence affecting urban areas, with projections indicating 19% of the world's population at risk by 2040.

Resource Extraction Activities

Underground extraction of solid minerals, such as coal, generates voids that induce subsidence through the collapse or sagging of overlying rock strata into the excavated space. This process is governed by mining method, extraction depth, seam thickness, and overburden lithology, with subsidence manifesting as gradual deformation or abrupt failures. In longwall mining, complete removal of coal panels promotes controlled caving, forming elliptical subsidence troughs where maximum vertical displacement reaches approximately 90% of the seam thickness for extraction widths exceeding 1.4 times the mining depth. For instance, at the York Canyon Mine near Raton, New Mexico, longwall operations at 107 meters depth in a 3-meter-thick seam produced up to 2.0 meters of subsidence beneath ridgetops. Room-and-pillar mining, which preserves support pillars for partial extraction ratios of 50-100%, typically yields delayed and irregular subsidence as pillars destabilize post-abandonment. Subsidence angles of draw extend 10°-35° beyond mine boundaries, influenced by rock strength; weaker shales facilitate earlier surface effects compared to competent sandstones. In Wyoming's Tongue River area, abandoned room-and-pillar coal mines under 15-30 meters overburden created subsidence pits and troughs severe enough to rupture a dam. Hydrocarbon extraction triggers subsidence via pore pressure depletion, which compacts reservoir sediments and transmits strain upward through elastic and plastic deformation. In the Wilmington Oil Field, Long Beach, California, production starting in 1932 caused up to 9 meters of subsidence across 22 square miles by the 1960s, with maximum bowl-center displacements exceeding 8 meters before mitigation via reinjection. Similarly, gas withdrawal from the Groningen field in the Netherlands has resulted in surface subsidence rates of about 9 millimeters per year in the field center over recent monitoring periods. Globally, oil and gas operations account for roughly 4% of documented land subsidence sites.

Surface Loading and Urban Development

Surface loading from urban development imposes additional mass on the Earth's crust, primarily through the construction of buildings, roads, and other infrastructure, leading to subsidence via consolidation of underlying compressible soils such as clays, silts, and peats. This process involves primary consolidation, where applied pressure expels pore water from saturated sediments, reducing volume, followed by secondary compression due to particle rearrangement and creep. In areas underlain by unconsolidated deltaic or alluvial deposits, this loading exacerbates subsidence, often interacting with but distinct from fluid withdrawal effects. Modeling studies quantify subsidence from urban loading as ranging from 5 to 80 mm in total, comprising elastic compression (0.2–3.2 mm), isostatic adjustment (0–20 mm), and nonlinear settlement (15–55 mm combined primary and secondary). These estimates derive from finite element simulations incorporating building footprints and soil properties, with nonlinear effects dominating in soft soils. In New York City, where total building mass reaches 7.64 × 10¹¹ kg, modeled subsidence in clay-rich fills and soils spans 75–600 mm (median 294 mm), though observed rates from InSAR and GPS data average 1–2 mm/year, suggesting ongoing consolidation. Specific cases illustrate localized impacts: in the San Francisco Bay region, with 1.6 × 10¹² kg of building mass supporting 7.75 million people, urban loading contributes to subsidence rates amplifying sea-level rise risks by 200–300 mm by 2050. In Hanoi, Vietnam's Red River Delta, InSAR monitoring from 2015–2021 reveals accelerated subsidence at over 40 new development sites on reclaimed agricultural land, where aggregate dumping and construction loading trigger rapid sediment consolidation beyond groundwater extraction alone. Similarly, in the urbanized coastal plain of the Netherlands, increased loading on peat soils via anthropogenic fill initially compresses subsurface layers, resulting in lower long-term subsidence rates (under 0.4 m) compared to adjacent agricultural areas (0.3–0.8 m), though future groundwater lowering could add 0.5–0.8 m in vulnerable zones. These examples highlight how urban expansion on weak foundations drives differential subsidence, posing hazards to infrastructure stability.

Measurement and Monitoring

Traditional Field Methods

Traditional field methods for subsidence monitoring rely on direct, ground-based measurements to quantify vertical displacements and deformations at specific points, offering high precision but limited spatial coverage compared to modern techniques. These approaches, developed primarily in the early 20th century, include spirit leveling, extensometers, and tiltmeters, which have been foundational in tracking subsidence in areas affected by groundwater extraction or mining. They emphasize empirical elevation changes tied to stable benchmarks, enabling repeated surveys to detect cumulative subsidence rates often exceeding 10-30 cm per year in vulnerable regions like California's San Joaquin Valley. Spirit leveling, one of the earliest and most accurate traditional methods, involves optical or digital levels and graduated rods to measure height differences along a network of benchmarks established by national geodetic surveys. Surveyors traverse lines between benchmarks, recording differential elevations with precisions down to 1-2 mm per kilometer of survey length, allowing detection of subsidence bowls or gradients over time. This technique has documented historical subsidence, such as over 9 meters in parts of the San Joaquin Valley since the 1920s, by comparing periodic re-levelings against fixed reference points. However, it requires extensive fieldwork, is susceptible to benchmark instability from subsidence itself, and provides only linear profiles rather than areal data. Extensometers provide subsurface measurements of compaction by installing vertical rods or wires anchored at depths below the subsiding zone, with surface plates tracking relative movement via dial gauges or transducers. Borehole extensometers, common in aquifer monitoring, can measure displacements to within 0.1 mm, isolating compaction in specific geologic layers as seen in extensometer arrays in Texas subsidence districts where annual rates reached 1-2 meters in the mid-20th century. These instruments directly link surface subsidence to irreversible aquifer skeleton compression, outperforming surface methods in resolving vertical strain profiles, though installation demands drilling and they capture only point-specific data. Tiltmeters complement these by detecting angular deformations indicative of differential subsidence, using spirit levels, pendulums, or electrolytic sensors to measure inclinations as small as 0.1 microradians. Deployed on structures or ground surfaces, they monitor tilt rates associated with subsidence troughs, as in mining areas where tilts exceed 10 microradians per month, providing early warnings of instability. Limitations include sensitivity to thermal effects and inability to distinguish vertical from horizontal components without integration with leveling data, necessitating calibration against benchmarks for absolute subsidence quantification. Overall, these methods remain essential for validating remote sensing in high-stakes sites due to their direct causality in linking observed motions to geomechanical processes.

Modern Remote Sensing Techniques

Interferometric Synthetic Aperture Radar (InSAR) represents the cornerstone of modern remote sensing for subsidence detection, leveraging phase interferometry from synthetic aperture radar (SAR) satellite imagery to measure centimeter- to millimeter-level ground displacements over areas spanning thousands of square kilometers. By comparing radar wave phases between repeat-pass acquisitions, InSAR quantifies line-of-sight surface changes with sub-millimeter precision under optimal conditions, enabling the identification of subsidence hotspots linked to groundwater extraction or tectonic activity. Satellites such as the European Space Agency's Sentinel-1 constellation, operational since April 2014 with 6- to 12-day revisit cycles and freely available C-band data, have facilitated widespread time-series monitoring, as demonstrated in studies of urban subsidence in regions like California's San Joaquin Valley, where rates exceeded 30 cm/year in some locales. Advanced InSAR variants enhance reliability in challenging environments: Persistent Scatterer InSAR (PS-InSAR) identifies stable radar reflectors, such as buildings or rocks, to mitigate decorrelation noise from vegetation or temporal changes, achieving deformation mapping with errors below 1 mm/year over multi-year spans. Small Baseline Subset (SBAS) techniques, by contrast, exploit multiple short-temporal-baseline interferograms to reduce atmospheric artifacts and unwrap phase ambiguities, as applied in monitoring subsidence from 2014 to 2022 in tectonically active basins. These methods have been validated against ground truth, with InSAR-derived rates correlating within 10-20% of extensometer measurements in overexploited aquifers, though limitations persist in densely vegetated or rapidly deforming terrains where phase unwrapping fails. Airborne and unmanned aerial vehicle (UAV)-based Light Detection and Ranging (LiDAR) complement satellite InSAR for high-resolution, local-scale subsidence assessment, generating digital elevation models (DEMs) with vertical accuracies of 5-10 cm to detect volumetric changes in mining-induced sinkholes or coastal erosion. Multitemporal UAV-LiDAR surveys, repeated at intervals of weeks to months, have quantified seasonal heave and subsidence in permafrost regions, revealing displacements up to several centimeters attributable to thaw cycles. Integration of LiDAR with InSAR and Global Navigation Satellite System (GNSS) receivers further refines measurements; for instance, GNSS-calibrated InSAR has reduced subsidence estimation errors to under 5 mm in integrated coastal studies, addressing InSAR's atmospheric delays through absolute positioning control. Such hybrid approaches, increasingly employed since the mid-2010s, enable comprehensive monitoring where single techniques falter, as evidenced by cross-validation yielding discrepancies below 0.6 cm/year in deltaic subsidence analyses.

Modeling and Prediction

Subsidence Modeling Approaches

Empirical models for subsidence prediction derive parameters from field observations and historical data, offering simplicity and speed for initial risk assessments, particularly in mining contexts where subsidence profiles are characterized by predictable patterns like troughs and offsets. These approaches, such as influence functions or profile functions developed for longwall coal mining, estimate maximum subsidence, width of influence, and offset based on extraction geometry and overburden properties, achieving reasonable accuracy in uniform geology but faltering in heterogeneous strata due to unaccounted variables like faulting or irregular extraction. For instance, empirical methods applied in New South Wales coal mines have demonstrated reliability for decision-making in stable conditions, though they require calibration and underperform compared to physics-based alternatives in variable lithologies. Numerical modeling employs finite difference, finite element, or meshless methods to solve coupled hydrogeological and geomechanical equations, capturing processes like poroelastic consolidation in aquifers or stress redistribution in mined voids. In groundwater-induced subsidence, models such as MODFLOW integrated with subsidence packages simulate aquitard compaction and interbed drainage, quantifying drawdown-subsidence relationships in compressible sediments; for example, applications in California's San Joaquin Valley have linked extraction rates to measured deformations exceeding 1 meter since the 1920s. Advanced implementations, including element-free Galerkin or ABAQUS-based simulations, incorporate nonlinear soil behavior and leakage through confining layers, enabling scenario testing for extraction policies but demanding high-quality geotechnical data and computational resources, with validation against extensometer or InSAR observations essential to mitigate parameterization errors. Data-driven approaches, leveraging machine learning, have gained traction for integrating heterogeneous datasets like InSAR time series, pumping records, and geospatial variables to forecast subsidence rates and susceptibility. Techniques such as long short-term memory (LSTM) networks or ensemble models like XGBoost predict temporal evolutions by correlating electricity consumption proxies for pumping with deformation trends, outperforming traditional regressions in nonlinear scenarios; a 2025 study in mining areas reported enhanced accuracy through SHAP interpretability for feature importance, though these models risk overfitting without physical constraints and require extensive training data. Hybrid frameworks combining numerical simulations with ML address uncertainties by assimilating real-time observations, as in Monte Carlo inversions for policy evaluation, yet demand rigorous cross-validation to ensure generalizability beyond site-specific calibrations.

Predictive Uncertainties and Limitations

Predicting land subsidence involves inherent uncertainties arising from the complex interplay of subsurface heterogeneity, incomplete data, and model simplifications. Hydrogeological parameters, such as aquifer compressibility and hydraulic conductivity, exhibit significant spatial variability that is challenging to characterize fully, leading to errors in model outputs that can exceed 20-50% in uncalibrated scenarios. Similarly, initial stress states and poroelastic properties in geomechanical models introduce parametric uncertainties, particularly in regions with limited borehole data, where assumptions about material behavior propagate through simulations. These factors are compounded by epistemic uncertainties from incomplete knowledge of fault systems or stratigraphic layers, which can alter predicted subsidence profiles nonlinearly. Structural uncertainties stem from the choice of modeling approaches, such as one-dimensional versus three-dimensional frameworks, which often fail to capture lateral flow or multi-aquifer interactions adequately. Empirical and semi-empirical models, reliant on historical data, perform poorly in novel geological settings or under changing extraction regimes, with prediction errors up to 30% reported in mining-induced cases due to unmodeled overburden dynamics. Physics-based numerical models, while more robust, demand high computational resources and struggle with long-term forecasts beyond 5-10 years, as irreversible compaction thresholds and feedback loops from surface loading are difficult to parameterize precisely. Data assimilation techniques, like ensemble Kalman filters, can reduce these by integrating InSAR observations, but residual uncertainties persist from measurement noise and temporal mismatches between satellite revisits (typically 6-12 days) and subsidence dynamics. Limitations in forecasting are further exacerbated by anthropogenic and climatic variables, including unregulated groundwater pumping or variable recharge rates, which defy deterministic modeling without real-time inputs. Global-scale predictions remain elusive due to sparse monitoring networks in developing regions, where subsidence rates may be underestimated by factors of 2-5 compared to localized studies. Machine learning alternatives, such as LSTM networks, improve short-term accuracy by capturing temporal patterns but inherit biases from training data scarcity and overlook causal mechanisms like clay consolidation, limiting their reliability for policy decisions. Overall, while uncertainty quantification via Monte Carlo methods or Bayesian inversion provides probabilistic bounds, predictions rarely achieve sub-centimeter precision over decadal horizons, underscoring the need for hybrid approaches integrating field validation.

Impacts

Environmental and Geomorphological Effects

Land subsidence compacts unconsolidated sediments and aquifers, causing irreversible lowering of the terrain and formation of earth fissures from differential settlement. These fissures rupture the surface, altering topographic gradients and disrupting natural drainage patterns, which can accelerate erosion or lead to anomalous sediment deposition. In regions like Arizona, subsidence rates of 1-2 inches per year have produced extensive fissure networks, modifying local landforms. In deltaic and coastal zones, subsidence amplifies relative sea-level rise, resulting in permanent inundation and wetland conversion to open water. For example, in Louisiana's coastal marshes, subsidence without sufficient sediment accretion has driven rapid marsh deterioration, with historical losses exceeding 100 km² annually in the 1980s due to combined subsidence and erosion processes. Similarly, the Indus River Delta has contracted to one-tenth its original extent from subsidence linked to groundwater overexploitation and reduced sediment input. Environmentally, aquifer compaction reduces permanent storage capacity, hindering recharge and promoting saltwater intrusion in coastal areas. In the Mekong Delta, subsidence at 1.6 cm per year has raised groundwater salinity to levels like 4.2 g/L, degrading water quality and stressing aquatic ecosystems. These changes heighten flood vulnerability, pond surface waters, and disrupt habitats, leading to biodiversity declines in subsidence-affected wetlands.

Infrastructure and Economic Consequences


Land subsidence induces structural damage to infrastructure through differential settling, which creates cracks, tilting, and misalignment in foundations, roads, bridges, pipelines, and utilities. In the United States, subsidence affects roads, gas and water lines, and building foundations, potentially leading to collapses and exacerbating coastal flooding. For instance, in California's San Joaquin Valley, where half the land is prone to sinking, subsidence has necessitated extensive repairs to canals and irrigation systems, with individual canal rebuilds costing up to $4.5 million as of 2015. Globally, maintenance costs for roads, railways, pipelines, and buildings rise significantly due to ongoing subsidence-induced stresses.
Economic consequences include direct repair expenditures, diminished property values, and indirect losses from disrupted services and increased insurance premiums. In China, anthropogenic subsidence results in average annual economic losses of approximately $1.5 billion, with 80-90% occurring in urban areas primarily from infrastructure damage. In California's Central Valley, areas experiencing subsidence lost 2.4% to 5.8% of home sale values due to groundwater depletion, translating to substantial dollar reductions in property markets. U.S. metropolises face heightened risks, with at least 20% of urban areas sinking—largely from groundwater extraction—affecting infrastructure for about 34 million people and amplifying long-term fiscal burdens. Additionally, subsidence alters flood zones, elevating property insurance costs and reducing land usability for development. In the Netherlands, subsidence damages building foundations, imposing high repair costs on homeowners amid rising public awareness of the issue.

Human and Flooding Risks

Land subsidence poses direct risks to human life primarily through sudden structural failures, such as building collapses or sinkhole formation, though such events are relatively rare compared to property damage due to the typically gradual nature of subsidence rates. In Lagos, Nigeria, subsidence exacerbated by groundwater extraction and poor construction has triggered over 200 building collapses since 2010, resulting in numerous fatalities and injuries. Globally, subsidence affects an estimated 19% of the world's population, or nearly 2 billion people, primarily through heightened vulnerability to infrastructure failure in urban areas. In the United States, subsidence impacts at least 34 million residents in major metropolises, where sinking land endangers transportation networks and housing stability. While peer-reviewed assessments emphasize that subsidence claims few lives directly—owing to warning signs from cracking structures—unmonitored anthropogenic drivers like aquifer overexploitation amplify hazards in densely populated regions. Subsidence significantly elevates flooding risks by reducing ground elevation, thereby increasing relative sea level rise and amplifying inundation during storms or high tides, often outpacing global sea level rise in affected coastal zones. Satellite interferometry data from 48 coastal cities worldwide reveal subsidence rates exceeding mean sea level rise (approximately 3-4 mm/year globally) in many locations, with sinking land contributing up to 10 times more to flood exposure than eustatic sea level changes in some deltas. For instance, in U.S. Atlantic and Gulf Coast cities, subsidence alone could submerge 11.9-15.1% of land below projected sea levels by 2050, independent of further ocean rise, affecting millions in low-lying infrastructure-heavy areas. Empirical modeling in regions like the Mississippi Delta shows subsidence compounding storm surge by 27-40% under historical hurricanes, deepening flood depths and expanding susceptible zones. This interplay heightens human exposure to flood-related casualties and displacement, particularly where subsidence masks or overshadows sea level narratives in risk assessments. In sinking megacities like Jakarta, annual subsidence rates of 10-15 cm have rendered 40% of the city below sea level, necessitating massive dike investments amid frequent inundation that displaces hundreds of thousands. Combined effects of subsidence and projected sea level rise (up to 74 cm by 2100 in vulnerability models) could increase coastal flood casualties by factors linked to elevated water levels, with historical data indicating 2.23 fatalities per meter of effective relative rise in prone areas. Such dynamics underscore subsidence's causal primacy in localized flooding over uniform global sea level trends, as evidenced by InSAR observations showing anthropogenic groundwater depletion as the dominant driver in 80% of high-subsidence hotspots.

Mitigation and Management

Engineering and Remedial Measures

Engineering measures for subsidence primarily address the underlying causes or stabilize affected structures, with remedial actions focusing on post-occurrence repair. For subsidence induced by groundwater extraction, artificial recharge of aquifers via injection wells or surface spreading basins replenishes water levels, reducing compaction in aquifer systems. In the Coachella Valley, California, substituting Colorado River and recycled water for groundwater pumping in projects since the 1970s has stabilized subsidence rates, demonstrating measurable recovery in land elevation. Repressurizing aquifers through dedicated wells similarly counters elastic and inelastic deformation, though effectiveness depends on aquifer permeability and prior irreversible compaction. In mining-related subsidence, backfilling mined-out voids with materials such as fly-ash cement slurries or granular aggregates prevents surface collapse by supporting overburden strata. High-volume grouting techniques, including gravity and compaction methods, fill fractures and voids, as applied in U.S. coal mine stabilization projects where slurries of up to 1:10 cement-to-water ratios achieved void filling with minimal surface disruption. Jet grouting and soil nailing reinforce weak zones, particularly in karst or abandoned mine settings, by creating soil-cement columns that enhance load-bearing capacity and limit differential settlement. Remedial structural interventions for buildings and infrastructure include underpinning, which extends foundations to stable strata using mini-piles or concrete segments, commonly employed in urban subsidence repairs to arrest further movement. Slab jacking lifts settled slabs by injecting cementitious or polyurethane grout beneath, restoring levelness with precision up to millimeters, though long-term durability varies with soil reactivity. Deep soil mixing creates stabilized ground columns in situ, mitigating ongoing subsidence in soft soils, as evidenced in projects combining mixing with recharge for comprehensive control. These measures often integrate with monitoring via extensometers or InSAR to verify efficacy, but challenges persist in retrofitting extensive infrastructure where costs can exceed millions per site.

Policy and Resource Management Strategies

In regions prone to subsidence from groundwater overexploitation, policies emphasize regulatory limits on extraction to maintain aquifer pressures and prevent irreversible compaction of fine-grained sediments. For instance, California's Sustainable Groundwater Management Act (SGMA) of 2014 requires local groundwater sustainability agencies to develop basin-specific plans that identify and mitigate subsidence risks, including thresholds for avoiding significant land sinking by sustaining groundwater levels above critical thresholds where compaction occurs. These plans integrate monitoring data from extensometers and satellite interferometry to track subsidence rates, often measured in inches per year, and prioritize actions like conjunctive use of surface and groundwater supplies to reduce pumping dependency. Complementing SGMA, California's Department of Water Resources issued draft Best Management Practices in July 2025 to guide local agencies in subsidence-prone areas such as the San Joaquin Valley, focusing on enhanced monitoring with new stations, site-specific extraction controls, and technical assistance to protect infrastructure like canals and levees from ongoing sinking. In Texas, subsidence districts such as the Harris-Galveston Subsidence District enforce groundwater regulation through permitting, production limits, and conversion to alternative supplies, achieving measurable reductions in subsidence rates since implementing data-driven plans in the 1970s; for example, the district mandates well-spacing rules and prioritizes surface water imports to curb overpumping in the Houston-Galveston region. Similarly, the Fort Bend Subsidence District regulates withdrawals to minimize subsidence contributions to flooding, emphasizing long-term shifts away from groundwater via infrastructure investments and conservation incentives. For subsidence induced by mining, resource management strategies incorporate pre-extraction assessments and post-mining liabilities. In the United Kingdom, the Coal Mining Subsidence Act 1991, administered by the Mining Remediation Authority, requires coal operators to predict subsidence via modeling, monitor surface impacts during extraction, and provide remedial works or compensation for damages, addressing historical liabilities from longwall mining that caused predictable surface troughs up to several meters deep. Australian guidelines for longwall coal mining mandate predictive subsidence simulations and real-time groundwater monitoring to mitigate hydrological disruptions, with regulatory approvals conditioned on minimizing surface and subsurface effects. Internationally, China's land subsidence control zones impose strict extraction quotas and supervision systems in overexploited deltas, combining enforcement with recharge projects to stabilize urban areas like Shanghai, where unmanaged pumping has exceeded 10 cm/year in localized spots. Resource allocation in these policies often involves economic instruments, such as fees on excess pumping or subsidies for recharge basins, to internalize the costs of permanent aquifer storage loss—estimated at billions in infrastructure repairs in cases like California's Central Valley, where subsidence has damaged canals requiring over $2 billion in fixes. Collaborative frameworks, including data-sharing among agencies and stakeholders, underpin adaptive management, though enforcement challenges persist in decentralized systems where local agricultural demands conflict with sustainability goals.

Economic Trade-offs and Criticisms

Managing groundwater extraction for agriculture exemplifies key economic trade-offs in subsidence, where short-term gains in crop production and farm profitability conflict with long-term costs from land sinking, including elevated pumping expenses, infrastructure damage, and heightened flood vulnerability. In California's Central Valley, excessive pumping to sustain irrigated farming—producing over $50 billion in annual agricultural value—has caused subsidence rates exceeding one foot per year in some areas since 2006, reducing nearby home values by up to 2-5% due to perceived risks and damaging canals and roads that require costly repairs. Globally, curbing overdraft to avert subsidence could diminish agricultural output by 0.73%, elevating staple food prices like rice by 7.4% and wheat by 6.7%, while increasing hunger risks for 26 million people by 2050, underscoring tensions between resource sustainability and food security in groundwater-reliant regions. Allocation policies further highlight these dynamics: "soft caps" on extraction, which allow flexible pumping limits, can boost expected farm profits by 12-13% over rigid "hard caps" by adapting to variable weather, yet they exacerbate aquifer depletion and associated subsidence by 12-13% cumulatively over decades, depending on aquifer transmissivity and well placement. Subsidence-induced economic damages amplify these trade-offs; in the U.S., average annual costs from related sinkhole and infrastructure failures exceed $300 million, with broader national repair burdens for highways, buildings, and utilities running into billions, often shifting taxpayer-funded mitigation onto extraction beneficiaries who capture private gains from resource use. Without intervention, subsidence could inflate global flood risks by $635 billion annually by 2050 through worsened coastal and urban vulnerabilities. Criticisms of subsidence management center on regulatory burdens that disproportionately affect extractive industries without commensurate alternatives, potentially stifling economic activity; for instance, California's groundwater restrictions under the Sustainable Groundwater Management Act have prompted fallowing of fields, contributing to $1.7 billion in drought-amplified revenue losses and 14,600 job cuts in agriculture by prioritizing aquifer recovery over immediate production needs. Cost-benefit analyses of mitigation, such as in Dutch urban areas like Gouda, reveal that strategies like foundation reinforcements or water injection yield mixed returns, with benefits-to-cost ratios below 1 in some scenarios due to high upfront investments and uncertain subsidence progression, questioning the efficiency of broad public subsidies for localized fixes. Moreover, policy inertia persists despite evident damages—exemplified by delayed responses in subsiding coastal cities—allowing cumulative losses to mount while extraction incentives remain, as governments hesitate to impose stringent limits that could disrupt rural economies in low- and middle-income countries reliant on overdraft for livelihoods. In regions like Indonesia's Semarang, public investments in subsidence countermeasures show positive net benefits only under optimistic assumptions, critiqued for overlooking opportunity costs in alternative infrastructure or agricultural adaptations. These approaches often externalize subsidence costs to non-extractors via elevated property taxes or insurance premiums, eroding incentives for private stewardship.

Notable Case Studies

Historical Instances

One of the earliest documented instances of subsidence occurred in the United Kingdom due to coal mining activities, with records of surface collapse dating back to the 18th century in regions like the Black Country and Northumberland. By the 19th century, widespread subsidence from room-and-pillar coal extraction had damaged infrastructure in industrial areas, including railways and buildings in Newcastle, where shallow mining led to predictable settlement patterns observed over decades. In Cheshire, brine extraction accompanying salt mining caused extensive sinkholes and subsidence troughs, with historical events documented from the mid-19th century onward, altering landscapes and prompting early remedial efforts like grouting. In Mexico City, subsidence was first noted in the late 19th century as the city, built on compressible lacustrine clays from the former Lake Texcoco, began sinking due to groundwater extraction for urban supply. By the early 1900s, annual subsidence rates reached approximately 5 cm in central areas, accelerating to peaks of 50 cm per year by mid-century from overexploitation of aquifers, causing differential settlement that tilted buildings and ruptured pipelines. Historical benchmarks, such as those installed in the 1920s, recorded cumulative drops exceeding 10 meters in some zones by 1950, with the phenomenon's onset linked directly to pumping volumes surpassing natural recharge. In the United States, the Goose Creek Oil Field near Baytown, Texas, marked one of the earliest documented cases of subsidence from fluid extraction, observed in the early 1900s when oil withdrawal caused surface drops of up to 2 meters by 1920. Similarly, in California's San Joaquin Valley, groundwater overdraft initiated significant subsidence in the mid-1920s, with initial pumping records from 1860 but measurable land-level declines first quantified around 1921, leading to over 9 meters of cumulative subsidence in parts of the valley by the 1970s. These events highlighted the causal link between aquifer compaction and prolonged extraction, damaging canals and aquifers irreversibly. Early 20th-century subsidence in Santa Clara Valley, California, provided another groundwater-driven example, with geodetic surveys from 1933 documenting initial drops attributed to agricultural pumping that intensified post-World War II. Across Europe and North America, such historical cases underscored subsidence as a consequence of resource extraction exceeding geomechanical limits, often without prior mitigation, leading to long-term landscape alterations.

Contemporary Examples

In Jakarta, Indonesia, excessive groundwater extraction to support rapid urbanization and industry has driven subsidence rates of up to 25 cm per year in northern coastal districts as of 2024, with some areas averaging 20 cm annually over the past decade. This compaction of aquifer sediments has lowered about 40% of the city's land below sea level by 2023, increasing flood vulnerability and prompting plans to relocate the capital to Nusantara. Regulatory restrictions on pumping since 2010 have slowed rates in regulated zones to under 5 cm per year, but illegal extraction persists in unregulated areas. Mexico City, built on the drained sediments of Lake Texcoco, experiences subsidence rates exceeding 40 cm per year in eastern boroughs like Iztapalapa due to overexploitation of aquifers supplying 70% of the city's water needs. Satellite interferometry data from 2015–2020 indicate maximum vertical displacements of 50 cm annually in high-extraction zones, causing differential sinking that damages infrastructure such as the metro system, with over 80 stations affected by track deformations and platform tilts. The city's total subsidence since the early 20th century exceeds 10 meters in places, but contemporary rates remain driven by unmet demand for 40 million cubic meters of groundwater daily, outpacing recharge. In California's San Joaquin Valley, groundwater overdraft during droughts, including 2012–2015 and post-2020 dry periods, has caused subsidence rates averaging 2.5 cm per year across large areas since 2006, with peaks up to 60 cm annually in Kern County hotspots. InSAR measurements through 2022 reveal cumulative sinking of over 1 meter in affected farmlands, reducing aquifer storage by billions of cubic meters and damaging canals like the California Aqueduct, which has required repeated repairs costing tens of millions. The Sustainable Groundwater Management Act of 2014 aims to curb pumping, yet enforcement delays have allowed ongoing elastic and inelastic compaction, with projections indicating potential infrastructure losses exceeding $2 billion if unchecked.
Along the U.S. Gulf Coast, such as in Houston and surrounding Texas counties, subsidence from oil, gas, and groundwater withdrawal has accelerated since the 2010s, with rates of 1–2 cm per year compounding storm surge risks in undetected zones identified via 2024 radar surveys. These patterns, linked to historical production exceeding 100 billion barrels of hydrocarbons, have deformed wetlands and pipelines, though precise quantification remains challenged by sparse monitoring.

Debates and Controversies

Attribution: Natural vs. Anthropogenic Factors

Natural subsidence arises from geological processes such as tectonic movements, glacial isostatic adjustment, and the autocompaction of deltaic sediments under their own weight. These mechanisms operate independently of human influence, with rates varying by region; for example, natural subsidence in the Venice Lagoon has been measured at approximately 0.9 mm per year, influenced by the heterogeneous subsoil composition. In tectonically active zones or post-glacial areas, such processes can dominate vertical land motion over millennia. Anthropogenic subsidence, conversely, stems primarily from subsurface fluid withdrawal, including groundwater pumping for agriculture and urban use, hydrocarbon extraction, and mining activities. In the United States, groundwater overuse accounts for about 80% of documented subsidence cases, particularly in California's San Joaquin Valley where historical rates exceeded 30 cm per year due to aquifer compaction. Globally, human-induced subsidence affects over 6.3 million km²—roughly 5% of land area—at rates above 5 mm per year, impacting nearly 2 billion people, with groundwater extraction as the leading driver in populated regions. Attributing subsidence to natural versus anthropogenic factors requires site-specific analysis, as the two often interact; for instance, human-engineered structures like levees can accelerate natural sediment compaction by preventing delta replenishment. In coastal settings, anthropogenic subsidence frequently amplifies relative sea level rise beyond global eustatic components, with fluid exploitation contributing dominantly where rates exceed 10 mm per year. Debates intensify here, as some assessments underemphasize subsidence's role—potentially due to institutional focus on climate narratives—leading to inflated attributions of flooding to sea level rise alone; empirical vertical land motion data, however, reveal subsidence as the primary local factor in many cases, such as U.S. Gulf Coast cities where it outpaces eustatic rise by factors of 2–5. Quantifying proportions remains challenging due to sparse monitoring and varying measurement techniques like InSAR and GPS, but peer-reviewed syntheses indicate anthropogenic drivers prevail in 70–90% of urban subsidence hotspots, while natural processes predominate in uninhabited or geologically dynamic areas. This distinction underscores causal realism: human interventions directly alter subsurface pressures, whereas natural subsidence reflects long-term equilibrium adjustments, with policy implications favoring targeted resource management over generalized climate framing.

Subsidence vs. Sea Level Rise in Flooding Narratives

In analyses of coastal flooding risks, relative sea level rise (RSLR)—the effective rise experienced locally—arises from the combination of global eustatic sea level rise (typically 3–4 mm per year in recent decades) and vertical land motion, including subsidence. In many coastal urban areas, subsidence rates exceed eustatic rise, dominating RSLR and accelerating flood exposure; for example, satellite interferometry data reveal subsidence surpassing sea level rise in numerous cities worldwide, with rates often linked to groundwater extraction and sediment compaction. This local sinking amplifies nuisance and storm flooding independently of global trends, as evidenced by projections showing subsidence alone could submerge significant land areas by 2050 in U.S. Atlantic and Gulf Coast cities, comprising 12–28% of below-sea-level zones without accounting for sea level contributions. Along the U.S. East Coast, subsidence rates average 1.3 mm per year (with maxima exceeding 10 mm per year), exposing 1.2–14 million people, 476,000–6.3 million properties, and over 50% of infrastructure in major metros like New York and Norfolk to heightened flood risks when combined with sea level rise. Such subsidence, driven by anthropogenic factors like fluid withdrawal rather than climatic forcing, can triple flooded areas over decades and increase nuisance flood frequency 3–12-fold by 2050 in affected regions. Globally, studies indicate subsidence contributes more to RSLR in coastal cities than prior estimates suggested, particularly in densely populated deltas where human activities exacerbate natural compaction. Flooding narratives, often centered on climate-driven sea level rise as the primary threat, have historically underemphasized subsidence's role, leading to projections that overlook its outsized local impacts. For instance, in regions like the U.S. Gulf Coast, subsidence from and gas extraction has historically accounted for much of observed RSLR, yet policy discussions prioritize emission reductions over site-specific interventions like aquifer recharge. Accounting for spatially variable subsidence refines risk models, revealing that unmitigated sinking could render 11.9–15.1% of Atlantic Coast land below by 2050 from subsidence alone, underscoring the need for integrated assessments beyond eustatic trends. This distinction highlights causal realism: while is uniform and gradual, subsidence is heterogeneous, reversible through management, and often the proximate cause of intensified flooding in vulnerable locales.

References

  1. [1]
    Land Subsidence | U.S. Geological Survey - USGS.gov
    Land subsidence is a gradual settling or sudden sinking of the Earth's surface due to removal or displacement of subsurface earth materials.
  2. [2]
    What is the difference between a sinkhole and land subsidence?
    Sinkholes are just one of many forms of ground collapse, or subsidence. Land subsidence is a gradual settling or sudden sinking of the Earth's surface.
  3. [3]
    Land Subsidence | U.S. Geological Survey - USGS.gov
    Land subsidence occurs when large amounts of groundwater have been withdrawn from certain types of rocks, such as fine-grained sediments. The rock compacts ...
  4. [4]
    What is subsidence? - NOAA's National Ocean Service
    Jun 16, 2024 · Subsidence can also be caused by natural events such as earthquakes, soil compaction, glacial isostatic adjustment, erosion, sinkhole formation, ...
  5. [5]
    Cause and Effect | U.S. Geological Survey - USGS.gov
    Although land subsidence caused by groundwater pumping has caused many negative effects on human civil works for centuries, especially in the highly developed ...
  6. [6]
    Subsidence | Regional Sea Level
    Subsidence is sinking land that leads to higher sea-level and increased flood risk. It is caused by processes like GIA, tectonics, and groundwater withdrawal.Missing: major | Show results with:major
  7. [7]
    Land subsidence risk to infrastructure in US metropolises - Nature
    May 8, 2025 · We estimate that at least 20% of the urban area is sinking in all cities, mainly due to groundwater extraction, affecting ~34 million people.
  8. [8]
    [PDF] Measuring Land Subsidence from Space - USGS.gov
    Land subsidence is a gradual settling or sudden sinking of the Earth's surface owing to subsurface movement of earth materials. Subsidence in the.
  9. [9]
    Land Subsidence in California | U.S. Geological Survey - USGS.gov
    Land subsidence is a gradual settling or sudden sinking of the Earth's surface due to subsurface movement of earth materials.Missing: definition | Show results with:definition
  10. [10]
    [PDF] 8 Types of land subsidence, by Alice S. Allen, Bureau of Mines, U.S. ...
    Land subsidence is a surface symptom of subsurface displacement, caused by processes that remove or rearrange subsurface materials to produce void space.Missing: definition | Show results with:definition
  11. [11]
    Land Subsidence in the U.S. (USGS Fact Sheet 165-00)
    Land subsidence is a gradual settling or sudden sinking of the Earth's surface due to subsurface movement, often caused by human impact on subsurface water.
  12. [12]
    Land subsidence due to withdrawal of fluids
    Land subsidence from fluid withdrawal is common in the US since 1940, caused by oil/gas and groundwater withdrawal, with examples in Texas, California, and ...<|separator|>
  13. [13]
    Subsidence - an overview | ScienceDirect Topics
    Subsidence is the sinking or lowering of a surface region relative to the surrounding area, due to the removal of material from underground formations.
  14. [14]
    [PDF] 3 Mechanics of land subsidence due to fluid withdrawal, by Joseph ...
    "The well-known hydrodynamic (Terzaghi) theory of soil consolidation can provide a semi- quantitative explanation for the phenomenon of repeated permanent ...
  15. [15]
    Insights into mechanisms of pumping-induced land subsidence ...
    Sep 17, 2025 · The mechanism of pumping-induced land subsidence is revealed from macro to micro. · The lagging land subsidence can be attributed to the creep of ...
  16. [16]
    Tectonic subsidence - LSU Center for GeoInformatics
    Three mechanisms are common in the tectonic environments in which subsidence occurs: extension, cooling and loading.
  17. [17]
    The Accommodation Space Equation - UGA Stratigraphy Lab
    Four main mechanisms that affect this isostatic balance and therefore drive tectonic subsidence include stretching, cooling, loading, and asthenospheric ...<|separator|>
  18. [18]
    Plate tectonics and basin subsidence history - GeoScienceWorld
    Tectonic setting controls basin formation. Passive margins have rapid initial and slow post-rift subsidence. Intracontinental basins have similar magnitude but ...
  19. [19]
    Subsidence across the Antler foreland of Montana and Idaho ...
    The complex patterns of subsidence across the Montana-Idaho foreland do not fit into simple flexural models for vertical loading of unbroken elastic plates.
  20. [20]
    Subsidence and Uplift (GH0309) - UNDRR
    Several crustal scale processes drive subsidence and uplift that tend to be regional or global in scale. Crustal movements occur in response to several ...
  21. [21]
    [PDF] Seismicity and Subsidence: Examples of Observed Geothermal ...
    Feb 24, 2014 · In 1987, the M 6.3 Edgecumbe earthquake, which was located at 8 km depth and about 20 km NE of Kawerau, caused significant co-seismic subsidence ...
  22. [22]
    Coseismic compression/dilatation and viscoelastic uplift/subsidence ...
    The 2012 Indian Ocean earthquake sequence (Mw 8.6, 8.2) is a rare example of great strike-slip earthquakes in an intraoceanic setting. With over a decade of ...
  23. [23]
    [PDF] 8 Subsidence
    In this chapter, the different mechanisms leading to crustal subsidence as well as some models quantifying the development of subsidence versus time, i.e., ...
  24. [24]
    Sedimentary basins: Regions of prolonged subsidence
    Jan 19, 2021 · An isostatic response (subsidence) as the mantle lithosphere adjusts to regional density contrasts, or,; Flexure (bending) of the lithosphere.
  25. [25]
    Quantifying subsidence and isostatic readjustment using ...
    Passive margins are characterised by an important tectonic and thermal subsidence, which favours a good preservation of sedimentary sequences.
  26. [26]
    U.S. East Coast Sea Level | Virginia Institute of Marine Science
    Glacial Isostatic Adjustment (GIA)​​ 2007). The maximum subsidence due to GIA (1.5 millimeters per year) is found around New Jersey (Karegar et al.
  27. [27]
    Land subsidence and relative sea-level rise in the southern ...
    Land subsidence has been observed since the 1940s in the southern Chesapeake Bay region at rates of 1.1 to 4.8 millimeters per year (mm/yr), and subsidence ...<|separator|>
  28. [28]
    Post-Glacial Isostatic Adjustment in Virginia | U.S. Geological Survey
    Sep 22, 2022 · Though it never extended as far south as Virginia, the ice sheet's historic presence still affects land subsidence rates in Virginia to this day ...
  29. [29]
    Subsidence - Tulane University
    Nov 3, 2016 · Carbonate rocks such as limestone, composed mostly of the mineral calcite (CaCO3) are very susceptible to dissolution by groundwater during the ...
  30. [30]
    Understanding sinkholes and karst - British Geological Survey
    The collapse may gradually propagate up through the overlying strata to cause subsidence at the surface. These sometimes extend up into rocks that are not ...
  31. [31]
    Karst, Sinkholes, and Land Subsidence | U.S. Geological Survey
    More than 80 percent of known land subsidence in the U.S. is a consequence of groundwater use, and is an often overlooked environmental consequence of our land ...
  32. [32]
    Sinkholes in Evaporite Rocks | American Scientist
    Surface subsidence can develop within a matter of days when highly soluble rocks dissolve because of either natural or human causes.
  33. [33]
    [PDF] Subsidence hazards in different types of karst
    The term "karst subsidence" refers to the surface features resulting from more or Jess Jong acting destructive processes, hidden in the subsurface, which ...
  34. [34]
  35. [35]
    Evaporite Karst Subsidence - Colorado Geological Survey
    It is this dissolution of the rock that creates caverns, open fissures, streams outletting from bedrock, breccia pipes, subsidence sags and depressions, and ...
  36. [36]
    [PDF] INTRODUCTION - Land subsidence in the United States
    The case histories illustrate the three basic mechanisms by which human influence on ground water causes land subsidence—compaction of aquifer systems, de-.
  37. [37]
    Global land subsidence mapping reveals widespread loss of aquifer ...
    Oct 4, 2023 · One of the most visible and harmful effects of groundwater depletion is land subsidence, which is caused by compaction of aquifer materials ...
  38. [38]
    Groundwater pumping and land subsidence in the Emilia‐Romagna ...
    Jan 19, 2006 · The results show that the extensive groundwater pumping that occurred in the past is most likely the main cause of the recent land settlement.<|separator|>
  39. [39]
    Groundwater pumping drives rapid sinking in California
    Nov 19, 2024 · A new study shows land in California's San Joaquin Valley has been sinking at record-breaking rates over the last two decades as groundwater extraction has ...
  40. [40]
    How 'sinking cities' can address subsidence challenges
    Oct 29, 2024 · Natural geological processes such as tectonic movements or seismic activity can also affect sediment compaction, contributing to subsidence. And ...
  41. [41]
    [PDF] Subsidence from Underground Mining: Environmental Analysis and ...
    Subsidence is a process from voids created by extracting solids/liquids, controlled by mining methods, depth, and deposit thickness, impacting surface ...
  42. [42]
    Subsidence Induced by Gas Extraction: A Data Assimilation ...
    Mar 20, 2022 · Surface movement can be induced by many human subsurface activities: production of natural gas, geothermal heat extraction, ground water ...<|separator|>
  43. [43]
    Reducing land subsidence in the Wilmington oil field by ... - OSTI.gov
    The subsidence at Long Beach attributed to the Wilmington oil field development encompassed an area of 22 sq miles where subsidence ranged from 2 ft to 30 ...
  44. [44]
    New subsidence map illustrates effect of Groningen gas production
    Nov 3, 2020 · Subsidence in the centre of the field amounts to around 9 mm/year over the past four years, whilst the areas surrounding the field show values ...
  45. [45]
    Global Land Subsidence: Impact of Climate Extremes and Human ...
    Nov 2, 2024 · Oil and gas extraction from underground formations has contributed to an estimated 4% of global LS sites surveyed, as shown in Figure 2e.4.1 Natural Resource... · 6.1 Peatlands And Oxidation... · 8 Research Gaps And...<|separator|>
  46. [46]
    The Weight of Cities: Urbanization Effects on Earth's Subsurface
    Jan 14, 2021 · The combined modeled subsidence from building loads is at least 5–80 mm, with the largest contributions coming from nonlinear settlement and ...
  47. [47]
    Urban development induced subsidence in deltaic environments
    Urbanisation drives subsidence in cities constructed on unconsolidated sediments. Spatial and temporal association with new developments and fast subsidence in ...
  48. [48]
    [PDF] The Weight of New York City: Possible Contributions to Subsidence ...
    Additional contributions to subsidence related to urbanization include pumping and/or other discharge of ground- water, and the rerouting of normal sediment ...
  49. [49]
    Differential subsidence in the urbanised coastal-deltaic plain of the ...
    Oct 9, 2018 · This form of subsidence is often human-induced, and is a consequence of phreatic groundwater level lowering and increased surface loading.Geological Setting And Study... · Peat Compression · Future Subsidence Estimates
  50. [50]
    Methods of Measuring Land Motion | U.S. Geological Survey
    Methods of Measuring Land Motion, including extensometers, Monitoring wells, InSAR, and GNSS Surveys
  51. [51]
    Measuring and Monitoring | U.S. Geological Survey - USGS.gov
    The most precise measurements tend to be made using spirit-leveling surveys and extensometers. The least precise measurements tend to be made by using GPS ...<|control11|><|separator|>
  52. [52]
    Measuring Subsidence - Harris Galveston Subsidence District
    Apr 14, 2025 · Traditional Surveying uses National Geodetic Survey (NGS) benchmarks to periodically document any changes to land surface.
  53. [53]
  54. [54]
    [PDF] Subsidence Management Monitoring Method
    The rate and extent of subsidence are the most useful parameters to measure. Several techniques may be used to monitor ground surface elevation including the ...
  55. [55]
    Methods of subsidence and sea level rise monitoring - USGS.gov
    Nov 16, 2022 · Five methods of monitoring subsidence and sea level rise: extensometers, InSAR, wells, GPS surveying, and tidal stations.
  56. [56]
    Tiltmeters monitor stability, differential settlement, and deformation
    Tiltmeters can trigger alarms when sudden movements occur, monitor structural stability, and monitor differential settlement and deformation.
  57. [57]
    Tiltmeters Provide Geotechnical Monitoring Solution | RST Instruments
    Tiltmeters allow for precise, up-to-the-minute monitoring of structures and can thereby provide an early warning of any shifts or changes.
  58. [58]
    Interferometric Synthetic Aperture Radar (InSAR) - USGS.gov
    Synthetic Aperture Radar (SAR) imagery is produced by reflecting radar signals off a target area and measuring the two-way travel time back to the satellite.
  59. [59]
    Review of satellite radar interferometry for subsidence analysis
    This paper includes a critical review of the existing literature on the use of satellite SAR imagery for subsidence analysis.
  60. [60]
    InSAR data for detection and modelling of overexploitation-induced ...
    Aug 2, 2024 · These limitations can be overcome with satellite InSAR techniques, a well-established approach for subsidence monitoring over recent decades, ...
  61. [61]
    Monitoring the Land Surface from Space
    Jul 23, 2024 · InSAR is a remote sensing technique that uses satellites to measure the movement of the land surface through the comparison of radar images collected at ...
  62. [62]
    Land subsidence analysis using synthetic aperture radar data
    The PS-InSAR approach used in this research is according to Ref. [32], which uses a single master and follows the amplitude dispersion criterion for pixel ...
  63. [63]
    Full article: Land subsidence monitoring using InSAR technique in ...
    Apr 15, 2024 · This study employed Interferometric Synthetic Aperture Radar analysis, spanning from 2014 to 2022, to effectively measure land subsidence, and visualize it ...
  64. [64]
    Ground subsidence monitoring in based on UAV-LiDAR technology
    Mar 1, 2024 · In this study, we propose a monitoring method for the ground subsidence of high-intensity mining with Unmanned Aerial Vehicle Lidar (UAV-LiDAR) measurement ...
  65. [65]
    Multitemporal UAV lidar detects seasonal heave and subsidence on ...
    Nov 26, 2024 · Our study shows the substantial value of using UAV lidar for understanding how permafrost areas are changing. It facilitates tracking the ...
  66. [66]
    An Integrated InSAR and GNSS Approach to Monitor Land ... - MDPI
    Nov 4, 2022 · The approach combines InSAR and GNSS, calibrating InSAR with GNSS, then integrating the data to monitor land subsidence, overcoming each ...<|control11|><|separator|>
  67. [67]
    Integrated coastal subsidence analysis using InSAR, LiDAR, and ...
    Dec 1, 2022 · This study is the first attempt to integrate satellite interferometric synthetic aperture radar (InSAR) and airborne light detection and ranging (LiDAR) methods
  68. [68]
    Analysis of regional large-gradient land subsidence in the Alto ...
    The statistics of the comparison between LiDAR and GNSS is minimum (Table 3). The discrepancy between LiDAR and GNSS can be observed that MaxD is 0.584 cm/year, ...
  69. [69]
    A critical review of mine subsidence prediction methods
    A review of the application of existing subsidence prediction methods available is presented. A brief description of all the three methods—empirical, functional ...
  70. [70]
    (PDF) Mining-Induced Subsidence Predicting and Monitoring
    Sep 17, 2025 · Empirical models remain effective in simple geological settings due to their low data and computational demands but lack adaptability in complex ...<|separator|>
  71. [71]
    Reliability of Subsidence Prediction Methods for Use in Mining ...
    There are several numerical models currently in use in NSW for predicting subsidence. The models do not appear to be superior in their ability to predict ...
  72. [72]
    A numerical model to calculate land subsidence, applied at Hangu ...
    A new computer code called the Interbed Drainage Package has been developed for the popular MODFLOW model to address this behaviour. It has been applied in a ...
  73. [73]
    Numerical modeling of land subsidence induced by groundwater ...
    This study aims to provide a coupled flow-deformation model for simulating land subsidence associated with groundwater extraction in aquifers.
  74. [74]
    A hydromechanical EFG-based Model for Numerical Simulation of ...
    Feb 17, 2024 · This study presents a coupled hydromechanical element-free Galerkin (EFG) model to simulate land subsidence induced by groundwater withdrawal.
  75. [75]
    ABAQUS-Based Numerical Analysis of Land Subsidence Induced ...
    This study investigates the developmental relationship between land subsidence and groundwater drawdown during pumping processes in complex leaky aquifer ...
  76. [76]
    Deep learning time-series modeling for assessing land subsidence ...
    Aug 22, 2025 · This study introduces a novel approach for analyzing and mitigating land subsidence by leveraging machine learning models and real-world ...
  77. [77]
    Machine learning-based techniques for land subsidence simulation ...
    Feb 14, 2024 · We introduce a comprehensive analysis of LS forecasting utilizing two advanced machine learning models: the eXtreme Gradient Boosting Regressor (XGBR) and Long ...
  78. [78]
    Land subsidence modelling for decision making on groundwater ...
    Apr 22, 2020 · This study presents an inversion scheme with uncertainty analysis for a land subsidence modelling by a Monte Carlo filter in order to contribute to the ...
  79. [79]
    A machine learning approach for mapping susceptibility to land ...
    In this study, we use the conventional Frequency Ratio (FR) method and ML models to generate LSSI maps of the region of Murcia (Spain) where land subsidence ...
  80. [80]
    Land Subsidence Model Inversion with the Estimation of Both Model ...
    Other Uncertainty Factors. This study primarily focused on model parameters and prediction uncertainties in land subsidence modeling. This section describes ...
  81. [81]
    Reducing uncertainty on land subsidence modeling prediction by a ...
    The main goal of this research was to develop a comprehensive geomechanical reservoir model that automatically and dynamically integrates the available ...
  82. [82]
    Geomechanics of subsidence above single and multi-seam coal ...
    The measured maximum subsidence above a single seam super-critical longwall panel (Smax) is typically 55%–65% of the seam's extracted thickness (T). The maximum ...Missing: trough | Show results with:trough
  83. [83]
    Unveiling the Global Extent of Land Subsidence: The Sinking Crisis
    Feb 20, 2024 · Understanding the subsidence process of a quaternary plain by combining geological and hydrogeological modelling with satellite InSAR data ...<|control11|><|separator|>
  84. [84]
    Land subsidence: A global challenge - PubMed
    Jul 15, 2021 · This study presents a comprehensive review of the Land subsidence (LS) cases, as a worldwide environmental, geological, and global geohazard concern.
  85. [85]
    a novel methodological approach for land subsidence prediction ...
    Apr 22, 2020 · The objective of this work is to propose a novel methodological approach for land subsidence prediction and uncertainty quantification.
  86. [86]
    Development and Comparison of InSAR-Based Land Subsidence ...
    Sep 9, 2024 · This study developed three time-series prediction models: a support vector regression (SVR), a Holt Exponential Smoothing (Holt) model, and multi-layer ...
  87. [87]
    Impacts of land subsidence caused by withdrawal of underground ...
    Jan 1, 2005 · Subsidence causes permanent inundation of land, aggravates flooding, changes topographic gradients, ruptures the land surface, and reduces the ...
  88. [88]
    The Fragile Fringe - Loss of Wetlands: Subsidence - USGS.gov
    Subsidence in the coastal marshes involves two factors: the sinking of the marsh surface and a lack of sediment being added to the marsh surface.
  89. [89]
    Delta Subsidence: An Imminent Threat to Coastal Populations
    Jul 31, 2015 · Apart from the increased risk of floods and associated diseases, experts consulted for this article say subsidence threatens health in other ...<|control11|><|separator|>
  90. [90]
    Study: From NYC to D.C. and beyond, cities on the East Coast are ...
    Jan 2, 2024 · Subsidence can undermine building foundations; damage roads, gas, and water lines; cause building collapse; and exacerbate coastal flooding – ...
  91. [91]
    Damage from sinking land costing California billions - CBS News
    Dec 27, 2015 · Rebuilding another canal recently cost $4.5 million. Putting a grand total on damage from subsidence in California is tricky because irrigation ...
  92. [92]
    1.4 Major Environmental Impacts – Land Subsidence and its Mitigation
    In China the average total economic loss due to anthropogenic land subsidence is estimated around 1.5 billion dollars per year 80‑90 percent of which are ...
  93. [93]
    Groundwater depletion sinks home prices in California's Central Valley
    or gradually sinking — areas lost between 2.4% and 5.8% of their sale value. In dollar terms, that ...
  94. [94]
    Policy Targeting to Reduce Economic Damages From Land ...
    Jun 4, 2018 · Land subsidence due to groundwater pumping generates economic costs by affecting flood zone designations and reducing property values ...
  95. [95]
    [PDF] Land subsidence exposed: The impact on house prices in the ...
    Nov 30, 2024 · In the Netherlands, subsidence damages houses and their foundations, resulting in high costs of repair for homeowners. However, public awareness ...<|separator|>
  96. [96]
    Land Subsidence Hazard and Building Collapse Risk in the Coastal ...
    Here, we highlight the role of land subsidence in triggering unprecedented collapses in the city of Lagos, Nigeria, which has reported over 200 casualties ...
  97. [97]
    Nearly 2 billion people globally at risk from land subsidence - Phys.org
    Mar 7, 2024 · Tahmasebi said, "Land subsidence is a destructive phenomenon that damages infrastructure and aquifers, as well as putting human lives at risk.Missing: casualties | Show results with:casualties
  98. [98]
    General Concept - UNESCO Land Subsidence International Initiative
    Subsidence is more hazardous to property than to life, because of the typically low rates of lowering. It has caused few casualties. Subsidence, however, ...
  99. [99]
    Subsidence in Coastal Cities Throughout the World Observed by ...
    Mar 24, 2022 · Satellite data indicate that land is subsiding faster than sea level is rising in many coastal cities throughout the world.
  100. [100]
    Disappearing cities on US coasts | Nature
    Mar 6, 2024 · Our analysis indicates that land areas below sea levels by 2050 resulting from only land subsidence account for 11.9–15.1% (Atlantic coast), ...
  101. [101]
    Projecting the effects of land subsidence and sea level rise on storm ...
    Relative to present day conditions, storm surge susceptible regions increase by 27% (Irene) to 40% (Matthew) due to subsidence. Combined with SLR (+ 74 cm), ...Sea Level Rise Projections · Results · Hurricane Irene
  102. [102]
    Scientists discover land subsidence may contribute more to coastal ...
    Dec 14, 2023 · Scientists discover land subsidence may contribute more to coastal flooding than previously thought. The new work combines methods to create ...
  103. [103]
    Land subsidence in Jakarta and Semarang Bay – The relationship ...
    Oct 1, 2021 · This paper introduces a combined natural and social science study to bring subsidence more to the forefront of coastal hazard research.
  104. [104]
    Human Lives at Risk because of Eustatic Sea Level Rise and ...
    Apr 1, 2015 · This weighted mortality rate is 2.23 fatalities per meter of raised sea level. b. Analysis of the effect of SLR on human lives. Equation (1) ...
  105. [105]
    [PDF] 7 Review of methods to control or arrest subsidence, by Joseph F ...
    Methods to control subsidence include reducing pumping draft, artificial recharge of aquifers from the land surface, and repressuring of aquifers through wells.
  106. [106]
    Mitigating land subsidence in the Coachella Valley, California, USA
    The implementation of three projects in particular – replacing groundwater extraction with surface water from the Colorado River and recycled water (Mid-Valley ...Missing: techniques ground
  107. [107]
    [PDF] SUBSIDENCE CONTROL BY HIGH VOLUME GROUTING
    The use of high volume grouting techniques utilizing flyash-cement grout slurries for mine subsidence control are discussed. A recent project is discussed to ...
  108. [108]
    Sinkholes: Engineering Solutions for Prevention and Remediation
    Techniques such as compaction grouting, jet grouting, and soil nailing are employed to reinforce the soil and reduce the risk of subsidence. Foundation Design: ...<|separator|>
  109. [109]
    Two ways to fight land subsidence | Opinion | Eco-Business
    May 22, 2023 · Artificial recharge and deep soil mixing are two techniques showing promise in rescuing sinking cities.
  110. [110]
    Subsidence - California Department of Water Resources
    Subsidence is the sinking of land due to soil or sediment changes, often caused by groundwater pumping, and is a growing issue in California.<|separator|>
  111. [111]
    DWR Releases Draft Best Management Practices on Managing ...
    Jul 24, 2025 · “Best Management Practices” Will Help Local Water Managers Combat Sinking Lands & Protect Critical Water Infrastructure. SACRAMENTO, Calif.
  112. [112]
    Groundwater Regulation - Harris Galveston Subsidence District
    Jul 9, 2025 · HGSD has successfully reduced subsidence rates by implementing a regulatory plan based on the best available groundwater and subsidence data, ...
  113. [113]
    Fort Bend Subsidence District
    Jun 2, 2025 · The District's purpose is to regulate the withdrawal of groundwater within the District to prevent subsidence that contributes to flooding and infrastructure ...FBSD Permitting Portal · District Management · Rules and Regulations · Meetings
  114. [114]
    Coal mining subsidence damage - a guide to your rights - GOV.UK
    Jan 21, 2019 · This guide describes your rights to claim under the 1991 Act if your home or other property is damaged by coal mining subsidence.Missing: management | Show results with:management
  115. [115]
    Monitoring and management of subsidence induced by longwall ...
    Nov 4, 2021 · This report describes modelling tools and approaches for predicting subsidence and monitoring and measuring the surface water and groundwater related impacts
  116. [116]
    Land Subsidence Control Zone and Policy for the Environmental ...
    A supervision system for the 2013-regulation was developed to prevent land subsidence, monitor the process, and mitigate geo-hazards caused by land subsidence ...
  117. [117]
    Sinking Lands, Damaged Infrastructure: Will Better Groundwater ...
    May 14, 2020 · Excess groundwater pumping can compact soils, causing land to sink. Because this subsidence can damage costly infrastructure, avoiding it is an important ...
  118. [118]
    Ending groundwater overdraft without affecting food security - Nature
    Jun 14, 2024 · Sustained GW overdraft also lowers aquifers' hydraulic head, raising pumping costs, causing land subsidence and saltwater intrusion, threatening ...
  119. [119]
    Hydrologic‐Economic Trade‐offs in Groundwater Allocation Policy ...
    Dec 11, 2020 · We examine the trade-offs of alternative groundwater allocation policies on farm profitability and stream depletion Multi-year allocations ...
  120. [120]
    How much does sinkhole damage cost each year in the United States?
    Sinkhole damages over the last 15 years cost on average at least $300 million per year. Since there is no national tracking of sinkhole damage costs, ...
  121. [121]
    A new framework for assessing the economic impact of land ...
    Feb 24, 2021 · Subsidence can have a major impact given increasing flood risk. It can even result in permanent land loss, damage to infrastructure and ...
  122. [122]
    Policy Brief: Drought and California's Agriculture
    Taking this into account, the drought's economic impact is estimated at $1.7 billion in revenue losses and 14,600 in lost jobs. In 2021, the drought hit ...
  123. [123]
    As CA cuts groundwater use, what happens to fallowed land?
    Apr 30, 2024 · With the state water board cracking down on groundwater, it is inevitable, experts say, that some fields will have to be taken out of production.<|separator|>
  124. [124]
    Cost-benefit analysis of urban subsidence mitigation strategies in ...
    Apr 22, 2020 · This paper presents the approach and outcomes of an exploratory cost-benefit analysis of subsidence mitigation strategies in the inner city ...
  125. [125]
    The 'wickedness' of governing land subsidence: Policy perspectives ...
    Jun 9, 2021 · We explore the paradox of slow political action in addressing subsiding land, particularly along high-density urban coastlines with empirical insights.
  126. [126]
    Cost-benefit analysis of mitigating subsidence damage in Semarang ...
    This study aims to analyze the economic rationale of alternative (public) investment strategies to reduce subsidence impact in the Semarang-Demak region.Missing: criticisms | Show results with:criticisms<|separator|>
  127. [127]
    Five decades of settlement and subsidence - GeoScienceWorld
    Oct 23, 2018 · Given the past mining history of these areas, subsidence problems were posed by the extraction of coal and other minerals (Taylor 1967; Price et ...
  128. [128]
  129. [129]
    [PDF] Unusual cases of mining subsidence from Great Britain, Germany ...
    Widespread subsidence due to the room and pillar mining and brine extraction of salt has been fairly well documented in parts of the UK, notably the Cheshire.
  130. [130]
    Over a Century of Sinking in Mexico City: No Hope for Significant ...
    Mar 30, 2021 · The overexploitation of the aquifer system is causing severe land subsidence in the ZMVM, which has been recognized since the early 1900s (Gayol ...
  131. [131]
    The Sinking of Mexico City
    In the early decades of this century, annual sinkage in the city center averaged about 2 inches, but when it peaked at mid-century the soil was collapsing away ...
  132. [132]
    Our History - Harris Galveston Subsidence District
    Jun 23, 2025 · Subsidence was first observed in the Houston region in the early 1900's. The Goose Creek Oil Field, near Baytown, TX, was the first documented case.Missing: famous | Show results with:famous
  133. [133]
    Land subsidence in the San Joaquin Valley, California, as of 1972
    Land subsidence which began in the mid-1920's due to groundwater overdraft in the San Joaquin Valley has caused widespread concern for the past two decades.Missing: famous | Show results with:famous
  134. [134]
    Selected Worldwide Cases of Land Subsidence Due to ... - MDPI
    The present review paper focuses on selected cases around the world of land subsidence phenomena caused by the overexploitation of aquifers.
  135. [135]
    1.2 Historical Framework – Land Subsidence and its Mitigation
    Rappleye (1933) provided the first specific records of subsidence due to groundwater pumping in the Santa Clara Valley (California, USA), and Ingerson (1941) ...
  136. [136]
    Current land subsidence in Jakarta: a multi-track SBAS InSAR ...
    Jul 2, 2024 · The significant subsidence in Jakarta is notably driven by groundwater extraction, particularly from the upper confined aquifer system. Urgent ...
  137. [137]
    How Jakarta has dug itself into a hole - 360info
    May 17, 2023 · In some places land subsidence is 20cm a year. Sea level rises are around 0.5cm/year. The rate of land subsidence is therefore far above that of ...
  138. [138]
    From Jakarta to Nusantara: Land subsidence and other pressing ...
    Aug 29, 2023 · Approximately 40% of Jakarta now lies below sea level as a result and predictive models suggest that the entire city will be underwater by 2050 ...
  139. [139]
    Jakarta - Coastal Defence Strategy and Flood Mapping - C40 Cities
    The restriction on groundwater extraction, decreed by regional regulation, is a key measure for slowing down land subsidence. The Northern Jakarta area is ...
  140. [140]
    The sinking of Mexico City | International - EL PAÍS English
    Sep 1, 2025 · The Valley of Mexico is subsiding at a rate of up to 30cm a year, a phenomenon exacerbated by the excessive extraction of groundwater, ...
  141. [141]
    Geohazard assessment of Mexico City's Metro system derived from ...
    Mar 12, 2024 · Land subsidence rates in Mexico City reach 500 mm/year, causing progressive damage to the city's core infrastructure, including the Metro system ...
  142. [142]
    Study identifies areas in Mexico City Metro affected by land ...
    Mar 20, 2024 · Excessive groundwater extraction in the city has led to rapid subsidence with rates reaching nearly 20 inches per year. Sharp changes in ...Missing: history | Show results with:history
  143. [143]
    Present-day land subsidence rates, surface faulting hazard and risk ...
    Highlights · Over 300 Sentinel-1 SAR scenes are used to monitor land subsidence in Mexico City. · InSAR reveals vertical displacement rates as high as −39 cm/year ...
  144. [144]
    Land Subsidence in the San Joaquin Valley | U.S. Geological Survey
    Reduced surface-water availability during 1976-77, 1986-92, 2007-09, and 2012-2015 caused groundwater-pumping increases in the San Joaquin Valley, declines in ...
  145. [145]
    Quantification of record-breaking subsidence in California's San ...
    Nov 19, 2024 · In California's San Joaquin Valley, groundwater overdraft has caused dramatic and continued land subsidence during two main periods, ...
  146. [146]
    Study finds land sinking at record pace in San Joaquin Valley
    Nov 25, 2024 · New research now shows that large portions of the San Joaquin Valley have sunk at a record pace since 2006.
  147. [147]
    Subtle Land Subsidence Elevates Future Storm Surge Risks Along ...
    Sep 3, 2024 · Our results reveal widespread subsidence features due to oil and gas production, groundwater pumping, and wetland degradation that were previously undetected.
  148. [148]
    Natural versus anthropogenic subsidence of Venice
    Sep 26, 2013 · A certain variability characterizes the natural subsidence (0.9 ± 0.7 mm/yr), mainly because of the heterogeneous nature and age of the lagoon subsoil.
  149. [149]
    Localized uplift, widespread subsidence, and implications for sea ...
    Sep 27, 2023 · Subsidence or uplift, referred to jointly as vertical land motion (VLM), can either diminish or enhance relative sea level rise (i.e., the rise ...
  150. [150]
    Subsidence critical factor in local sea level rise assessments
    Subsidence is caused by both human activities (anthropogenic subsidence) and natural processes. Anthropogenic Subsidence has many causes such as:.
  151. [151]
    Land Subsidence | NASA Earthdata
    Land subsidence can lead to sink holes and infrastructure damage, aquifer-system compaction, subsequent reduced groundwater availability, and other issues. Land ...
  152. [152]
    Often driven by human activity, subsidence is a problem worldwide
    May 12, 2021 · Almost one-fifth of the planet's population lives in areas where subsidence driven by groundwater withdrawals is a major threat, a new analysis finds.
  153. [153]
    Coastal subsidence and relative sea level rise - IOPscience
    Sep 23, 2014 · Anthropogenic subsidence may be the dominant contributor to relative sea-level rise in coastal environments where subsurface fluids are heavily exploited.
  154. [154]
    Coastal Inundation: Rising Sea Levels Explained
    Mar 13, 2024 · When coastal subsidence and uplift are taken into account, claims of extreme rates of increases in sea level rise appear to be incorrect.
  155. [155]
    Climate Change: Global Sea Level
    Global average sea level has risen 8-9 inches since 1880, and the rate is accelerating thanks to glacier and ice sheet melt.Highlights · Why Sea Level Matters · Future Sea Level Rise<|control11|><|separator|>
  156. [156]
    Exposure of communities and infrastructure to subsidence on the US ...
    Jan 2, 2024 · The maximum subsidence rate across the US east coast exceeds 10 mm per year (Fig. ... glacial isostatic adjustment model” by Purcell et al.
  157. [157]
    Subsidence coastal cities contributes more to relative sea level rise ...
    In parts of these coastal cities, local land subsidence due to, in particular, human activities dominates over absolute sea level rise and land movement ...
  158. [158]
    Global Sea-Level Rise and Subsidence - Physical Agents of Land ...
    Global Sea-Level Rise and Subsidence. Submergence refers to permanent flooding of the coast caused by a rise in global sea level and/or subsidence of the land.