Fact-checked by Grok 2 weeks ago

Anderson's theory of faulting

Anderson's theory of faulting is a seminal framework in that explains the initiation, orientation, and mechanics of faults in the brittle upper crust as a function of the regional stress field. Developed by Scottish geologist Ernest Masson Anderson, the theory posits that fault planes form at optimal angles to the principal stress directions to accommodate shear failure under varying tectonic conditions. The theory rests on two key assumptions applicable to shallow crustal levels: the Earth's surface represents a free boundary with no , implying that the principal stresses align with the vertical and horizontal directions; and the vertical stress arises primarily from the lithostatic load of the overlying rock column. These assumptions lead to three fundamental tectonic regimes defined by the relative magnitudes of the principal stresses (denoted σ₁ as the maximum compressive, σ₂ intermediate, and σ₃ minimum): the gravity regime where σ₁ is vertical, favoring extensional faulting; the thrust regime where σ₃ is vertical, promoting compressional reverse faulting; and the wrench regime where σ₂ is vertical, resulting in strike-slip faulting. Building on the Coulomb-Mohr failure criterion, Anderson's model predicts conjugate fault pairs inclined at approximately 30°–45° to the maximum direction, yielding characteristic dips of about 60° for faults, 30° for low-angle reverse () faults, and near-vertical planes for strike-slip faults, assuming a typical rock of 30°. First outlined in Anderson's 1905 address to the Edinburgh Geological Society and elaborated in his 1942 monograph The Dynamics of Faulting and Dyke Formation with Applications to , the theory has profoundly influenced modern , , and by providing a basis for inferring paleostress orientations from fault geometries observed in the field.

Introduction

Historical development

Ernest Masson Anderson, a Scottish trained at the , first proposed his theory of faulting in 1905 during a to the Edinburgh Geological Society. In his paper "The dynamics of faulting," Anderson introduced a classification of faults based on the orientation of the three principal es, assuming that one principal stress acts vertically due to the influence of the Earth's free surface. This work marked a pioneering application of stress analysis to geological structures, predating widespread recognition of strike-slip faulting, as evidenced by the . Anderson's ideas drew on late 19th-century foundations in . His development occurred in the context of early 20th-century advances, including his appointment to the Geological Survey of in 1903, which provided field opportunities in to observe fault patterns, and growing interest in tectonic processes following continental drift hypotheses. These influences allowed Anderson to integrate with empirical observations, establishing a theoretical basis for fault under varying regimes. Anderson expanded and formalized his framework in the book The Dynamics of Faulting and Dyke Formation with Applications to Britain, first published in 1942 and revised in a second edition in 1951. This seminal text elaborated on the mechanics of faulting, incorporating the failure criterion and applications to , solidifying the theory's role in . The centennial of Anderson's 1905 paper was marked by reflections on its lasting influence, particularly in a 2008 review that analyzed global seismicity data to affirm the prevalence of Andersonian fault types in focal mechanisms. This recognition underscored the theory's foundational status, with over a century of refinements in and continuing to build upon its core assumptions.

Core principles

Anderson's theory of faulting establishes that shear in the brittle crust occurs on planes inclined at approximately ° to 45° relative to the maximum principal direction (σ₁), positioning the fault roughly perpendicular to the minimum principal stress (σ₃). This minimizes the shear required for while maximizing resistance due to normal stress on the plane, assuming frictional sliding as the dominant mechanism. The specific angle depends on the internal angle of , typically around ° for common crustal materials with a of about 0.6. Central to the theory is the assumption that the Earth's imposes a boundary condition where no acts on horizontal planes, making the vertical direction one of the principal axes. Consequently, the vertical (σᵥ) aligns with either σ₁, σ₂, or σ₃, simplifying the of fault orientations in a three-dimensional field. This vertical principal arises from the and gravitational loading, influencing how horizontal tectonic forces interact with the crust. The theory classifies faults into three primary types based on which principal stress is vertical, directly linking stress regimes to fault geometry:
  • Normal faults, formed when σ₁ is vertical (overburden dominates) and σ₃ is horizontal, accommodating crustal extension with typical dips of around 60°.
  • Reverse and thrust faults, occurring when σ₃ is vertical and σ₁ is , reflecting horizontal contraction with shallower dips of about 30°.
  • Strike-slip faults, developing when σ₂ is vertical and both σ₁ and σ₃ are , producing near-vertical fault planes for lateral shear.
Conjugate fault pairs represent a key prediction, consisting of two symmetric sets of faults mirrored across the principal axes, enabling balanced shear displacement in opposite senses while maintaining overall equilibrium. These pairs form at angles of 45° ± (friction angle)/2 to σ₁, providing insight into the bisected nature of the failure envelope.

Stress Fundamentals

Stress and strain in rocks

Stress is defined as the force applied per unit area on a rock body, quantified in units such as or . In the , rocks experience two primary categories of : lithostatic , which arises from the uniform of overlying rock layers and acts isotropically in all directions, and tectonic , which results from plate movements and is typically anisotropic, leading to differential loading. Lithostatic increases linearly with depth due to the weight of the rock column above, while tectonic superimposes directional components that drive deformation. Rocks are subjected to three fundamental types of based on the direction of force application: , which shortens the rock; tensile stress, which elongates it; and , which causes sliding or distortion parallel to a . These can be decomposed into hydrostatic (or mean) , representing the isotropic component equal in all directions, and deviatoric stress, capturing the differential part that promotes shape change without volume alteration. In crustal settings, hydrostatic stress dominates under lithostatic conditions, but deviatoric tectonic is crucial for initiating faulting by amplifying shear components on potential failure planes. Strain represents the response of a rock to applied , manifesting as a change in volume or shape, often expressed as a dimensionless . Rocks exhibit at low levels, where deformation is reversible and proportional to , following ; plastic (ductile) involves permanent deformation through mechanisms like crystal plasticity; and brittle failure occurs via fracturing when exceeds strength. The dominant response varies with environmental conditions: brittle failure prevails at shallow depths (<10-15 km) and low temperatures (<300-400°C) under high rates, transitioning to plastic flow at greater depths and higher temperatures where pressure suppresses fracturing. This transition influences fault susceptibility, as brittle regimes favor discrete fault formation. Key rock properties governing fault susceptibility include cohesion, the inherent tensile strength resisting initial failure, and the internal friction angle, which quantifies resistance to shear sliding along planes, typically ranging from 20° to 40° for crustal rocks depending on mineralogy and texture. Higher cohesion enhances resistance to brittle failure, while lower friction angles, often seen in clay-rich or weathered rocks, promote easier slip and fault reactivation. These parameters, derived from laboratory triaxial tests, are integral to assessing how rocks deform under combined stress regimes. To evaluate stress orientation on arbitrary planes within a rock, the normal stress \sigma_n and shear stress \tau on a plane inclined at angle \theta to the maximum principal stress \sigma_1 (with minimum principal stress \sigma_3) are given by: \sigma_n = \frac{\sigma_1 + \sigma_3}{2} + \frac{\sigma_1 - \sigma_3}{2} \cos(2\theta) \tau = \frac{\sigma_1 - \sigma_3}{2} \sin(2\theta) These equations highlight how shear stress maximizes at \theta = 45^\circ, providing a foundation for understanding potential fault orientations under differential loading.

Principal stresses and Mohr circle

In rock mechanics, principal stresses are defined as the maximum (σ₁), intermediate (σ₂), and minimum (σ₃) normal stresses acting on mutually perpendicular planes where shear stress vanishes, with the convention σ₁ ≥ σ₂ ≥ σ₃ assuming compression positive. These stresses represent the eigenvalues of the three-dimensional stress tensor, which encapsulates the full state of stress at a point in a continuum. The corresponding principal directions are the eigenvectors of this tensor, determined qualitatively through diagonalization—a process that transforms the coordinate system to align with axes where the tensor becomes diagonal, eliminating off-diagonal shear components. In the geological context of faulting, σ₁ denotes the axis of maximum compressive stress, while σ₃ indicates the axis of minimum compressive stress (or maximum extension), with σ₂ being the intermediate value. This orientation framework is fundamental to analyzing how applied stresses lead to deformation in the Earth's crust, as the principal axes dictate the planes of pure normal loading without superimposed shear. The Mohr circle offers a graphical tool to visualize and compute the stress components on arbitrary planes within a given stress state. For a two-dimensional analysis in the plane normal to σ₂, it is plotted with normal stress σ_n on the abscissa and shear stress τ on the ordinate; the circle's center lies at the mean stress \frac{\sigma_1 + \sigma_3}{2}, and its radius is \frac{\sigma_1 - \sigma_3}{2}, yielding a diameter equal to the differential stress σ₁ - σ₃. \begin{aligned} \text{Center: } & \quad \left( \frac{\sigma_1 + \sigma_3}{2}, 0 \right) \\ \text{Radius: } & \quad \frac{\sigma_1 - \sigma_3}{2} \end{aligned} Points on the circle represent the σ_n and τ for planes at various orientations, where the angular position on the circle is twice the physical angle from the principal direction due to the quadratic nature of stress transformations. This construction facilitates rapid determination of stresses on any plane and highlights the principal planes as the intercepts on the σ_n axis, where τ = 0. In three dimensions, multiple Mohr circles describe the full stress state, with the largest circle (between σ₁ and σ₃) governing the maximum shear stresses relevant to deformation analysis.

Theoretical Framework

Key assumptions

Anderson's theory of faulting relies on a set of simplifying assumptions to predict the orientation and type of faults based on the tectonic stress regime. These premises idealize the brittle upper crust as a medium where failure occurs predictably under differential stress, drawing from the while adapting it to geological conditions. By assuming a uniform stress field and specific boundary conditions, the model enables the classification of normal, reverse, and strike-slip faults without accounting for complications like fluid effects or material heterogeneities in its foundational form. A central assumption is that one of the principal stresses—one of \sigma_1 (maximum), \sigma_2 (intermediate), or \sigma_3 (minimum)—is oriented vertically, arising from the traction-free nature of the Earth's surface, which imposes no shear traction at shallow depths. This vertical principal stress aligns with the direction of gravity and lithostatic pressure, simplifying the three-dimensional stress tensor to cases where the overburden stress is one of the extrema. The theory further posits that faults initiate as shear fractures oriented at the angle dictated by the rock's internal friction, commonly taken as approximately 30° relative to the maximum principal stress \sigma_1. This orientation emerges from the geometry of shear failure in brittle materials, where the fracture plane bisects the angle between \sigma_1 and \sigma_3 adjusted for frictional resistance. Additionally, the model assumes a homogeneous and isotropic brittle crust with a uniform stress field at the depths where faulting occurs, treating the rock mass as mechanically uniform without significant variations in strength or preexisting weaknesses. Initially, the framework neglects the effects of pore fluid pressure and material anisotropy, focusing on effective stress in dry or undrained conditions, although Anderson later acknowledged their roles in modifying fault behavior. These assumptions collectively imply characteristic fault dips: approximately 60° for normal faults (where \sigma_1 is vertical and \sigma_3 is horizontal), 30° for reverse faults (where \sigma_1 is horizontal and \sigma_3 is vertical), and near-vertical (90°) planes for strike-slip faults (where \sigma_2 is vertical).

Coulomb failure criterion and fault initiation

The Coulomb failure criterion describes the conditions under which rocks undergo shear failure, forming the mechanical basis for fault initiation in Anderson's theory. This criterion states that failure occurs on a plane when the shear stress \tau acting parallel to it exceeds the rock's shear strength, expressed as \tau = C + \mu \sigma_n, where C is the cohesion (intrinsic tensile strength), \mu is the coefficient of internal friction, and \sigma_n is the effective normal stress perpendicular to the plane. In rocks, \mu typically ranges from 0.6 to 0.85, equivalent to a friction angle \phi = \tan^{-1} \mu \approx 30^\circ to 40^\circ, reflecting the frictional resistance derived from mineral grain interactions and surface roughness. Under a general triaxial stress state with principal stresses \sigma_1 > \sigma_2 > \sigma_3 (where \sigma_1 is the maximum ), failure initiates on the where the \tau / \sigma_n is maximized relative to the material's strength. This optimal is oriented at an angle \theta = 45^\circ - \phi/2 to \sigma_1, ensuring the is sufficient to overcome frictional locking while minimizing the stabilizing normal stress. For \phi \approx ^\circ, \theta \approx ^\circ, meaning incipient faults form at shallow angles to the direction of maximum compression. The Mohr circle representation illustrates this by showing how the stress trajectory tangentially intersects the linear envelope defined by the criterion. In Anderson's framework, this integrates with the that one principal (\sigma_v, the vertical lithostatic ) remains fixed by gravitational loading near Earth's surface, constraining the possible orientations and types of faulting. Conjugate shear pairs emerge symmetrically about \sigma_1, separated by 2\theta \approx 60^\circ, as both planes reach the threshold simultaneously under conditions. The vertical \sigma_v then dictates the tectonic : if \sigma_v = \sigma_1 (maximum), extensional normal faulting with \sigma_3 (minimum); if \sigma_v = \sigma_3 (minimum), compressional reverse faulting with \sigma_1 (maximum); if \sigma_v = \sigma_2 (), strike-slip faulting with \sigma_1 and \sigma_3. This setup predicts that fault requires the components to build until the threshold is met on the conjugate planes aligned with the . Faults begin as Mode II (in-plane ) fractures, where displacement is parallel to the fracture surface and perpendicular to the propagation direction, consistent with the mechanism under differential compression. As slip accumulates, these initial fractures propagate and evolve, potentially developing listric geometries (concave-upward curvature) in ductile layers or remaining planar in brittle ones, influenced by rock and evolving stress fields. The critical differential stress required for failure under the Coulomb criterion, assuming the optimal plane orientation, is \sigma_1 - \sigma_3 = \frac{2C \cos \phi}{1 - \sin \phi} + \frac{2 \sigma_3 \sin \phi}{1 - \sin \phi}. In the simplified cohesionless case (C = 0), this reduces to \sigma_1 - \sigma_3 = \sigma_3 \frac{2 \sin \phi}{1 - \sin \phi}, or equivalently \frac{\sigma_1}{\sigma_3} = \frac{1 + \sin \phi}{1 - \sin \phi}, highlighting the role of confining \sigma_3 in enabling frictional sliding. For \phi \approx ^\circ, the ratio is approximately 3, meaning \sigma_1 must exceed \sigma_3 by at least threefold under low cohesion to initiate faults.

Fault Classifications

Normal faults

In Anderson's theory of faulting, normal faults form under an extensional stress regime where the maximum principal (σ₁) acts vertically due to lithostatic load, while the minimum principal (σ₃) is oriented horizontally in the direction of extension, and the intermediate stress (σ₂) is the other horizontal direction. This configuration, often termed the "gravity regime," promotes failure along planes inclined at approximately 60° to the horizontal near the surface, as predicted by the failure criterion assuming typical rock angles around 30°. The theory posits that such faults initiate as planar shear fractures but may evolve into listric geometries that flatten with depth, accommodating greater displacement in the subsurface. Geometrically, normal faults exhibit a hanging wall that moves downward relative to the footwall, creating characteristic structures such as systems. Conjugate pairs of normal faults typically form symmetrically about the σ₁ direction (vertical), with antithetic faults dipping opposite to the main synthetic set, both striking to the σ₃ direction to maximize resolution. This symmetry arises from the theory's assumption of a homogeneous, isotropic medium under uniform stress, leading to fault planes oriented at about 30° to σ₁ for optimal failure. Representative examples of normal faulting aligned with Anderson's model include the in the , where high-angle normal faults accommodate widespread horizontal extension and crustal thinning over millions of years. Similarly, rift zones like the demonstrate this regime, with fault arrays forming half-grabens and promoting continental breakup. Kinematic indicators on normal faults include down-dip slickensides, which record the direction of hanging wall downdrop, and fault-plane solutions that reveal extensional mechanisms with P-axes aligned vertically (parallel to σ₁) and T-axes horizontal (parallel to σ₃). These features confirm the slip vector's consistency with the predicted field. The predicts that normal faults will strike perpendicular to the σ₃ direction, effectively accommodating horizontal stretching and vertical thinning of the crust without significant strike-slip component. This orientation facilitates efficient extensional deformation, as seen in regions where fault reactivation aligns with ongoing tectonic extension.

Reverse and thrust faults

In Anderson's theory of faulting, reverse and thrust faults develop in a compressional tectonic regime where the minimum principal (σ₃) is oriented vertically, governed primarily by the lithostatic , while the maximum principal (σ₁) acts horizontally in the of , and the intermediate (σ₂) is the other horizontal . This configuration, combined with the failure criterion and a typical internal (φ) of approximately 30° for rocks, predicts that optimally oriented conjugate fault planes will dip at about 30° to the horizontal. The theory assumes a homogeneous, isotropic brittle crust near the surface, where the minimum principal aligns with (vertical), combined with horizontal tectonic making the maximum principal horizontal. Geometrically, these faults exhibit reverse motion, with the hanging wall displaced upward relative to the footwall, and can manifest as low-angle thrusts (dips <30°) or steeper high-angle reverse faults (dips up to 50°–70°), the latter often representing reactivated or non-optimal planes. In mature compressional settings, such as fold-thrust belts, reverse and thrust faults commonly arrange into imbricate fans—overlapping series of subparallel thrusts that stack crustal slices and promote wedge-shaped thickening of the sedimentary cover. Kinematic indicators include slickenlines raking near 90° on the fault plane, indicating pure dip-slip motion, as well as overturned strata and asymmetric folds adjacent to the fault, where the hanging wall sequences are inverted due to the upward transport. Prominent examples of Andersonian reverse and thrust faulting occur in foreland basins and orogenic wedges, such as the fold-thrust belt of the European Alps, where Miocene compression produced imbricate thrusts like the Pennine Front, accommodating Africa-Europe convergence through low-angle detachment and ramping structures. Similarly, the Appalachian Valley and Ridge province features thrust sheets that align with the predicted 30° dips in shallow crustal levels, illustrating the theory's application to continental collision zones. While the pure Andersonian model assumes constant fault orientation, real-world thrusts often exhibit listric geometry, flattening at depth due to increasing overburden and pore pressure effects, which reduce effective normal stress and allow deviations from the ideal 30° dip toward even lower angles (e.g., 15°–20°). This adaptation arises from non-hydrostatic stress gradients and anisotropic layering, slightly modifying the theory's predictions without invalidating its foundational principles for brittle fault initiation.

Strike-slip faults

In Anderson's theory of faulting, strike-slip faults develop in a wrench tectonic regime where the intermediate principal stress (σ₂) is vertical due to the influence of lithostatic pressure near Earth's surface, while the maximum (σ₁) and minimum (σ₃) principal stresses are horizontal and oriented at approximately 45° to the fault plane. This configuration arises from the , which predicts shear failure along planes optimally oriented to the principal stress directions under these conditions. The geometry of strike-slip faults in this regime consists of near-vertical fault planes that accommodate purely horizontal relative displacement between the hanging wall and footwall, resulting in either right-lateral (dextral) or left-lateral (sinistral) motion depending on the sense of shear. In regions of restraining bends along these faults, where the fault trace curves to impede lateral motion, positive flower structures can form, characterized by upward-splaying subsidiary faults that root into a subvertical master fault and produce localized uplift. Prominent examples include the San Andreas Fault system in California, a right-lateral strike-slip fault that exemplifies Anderson's predictions for transform boundaries between tectonic plates, with horizontal slip rates averaging about 3–5 cm per year. Kinematic indicators on strike-slip faults typically include horizontal slickensides—polished and striated fault surfaces with lineations parallel to the slip direction—and en echelon arrays of subsidiary shears or tension gashes that form at low angles to the main fault trace, revealing the sense of lateral movement. The theory predicts that strike-slip faults will strike parallel to the σ₂ direction, maximizing shear stress on vertical planes while producing no net dip-slip component, though conjugate shear pairs may develop at acute angles to σ₁ for balanced accommodation of horizontal strain.

Oblique faults

Oblique faults, also known as oblique-slip faults, exhibit hybrid movement combining both dip-slip and strike-slip components, resulting in slip vectors that are neither purely vertical nor horizontal. These faults represent an extension of Anderson's classical theory of faulting, which primarily predicts pure normal, reverse, or strike-slip motions under the assumption that one principal stress axis is vertical due to gravitational loading near the Earth's surface. By relaxing this assumption—allowing the maximum (σ₁) and minimum (σ₃) principal stresses to be inclined relative to the vertical—oblique faulting emerges as a natural outcome, particularly in regions where tectonic forces deviate from simple vertical-horizontal alignments. This inclination leads to non-optimal fault orientations, where slip occurs on planes that accommodate both shear components to minimize overall stress. In the stress regime for oblique faults, the obliquity of σ₁ and σ₃ to the surface produces faults with variable dips, typically moderate (around 45–60°), and rakes (slip angles) that deviate significantly from 0° (pure ) or 90° (pure ). Slickenside lineations on fault surfaces often trend obliquely, indicating the combined motion, and fault planes may form en echelon arrays to better align with the resolved shear stress. Such configurations arise because the rotated principal stresses cause the Coulomb failure criterion to favor hybrid slip directions on pre-existing or newly formed fractures, rather than the conjugate pairs predicted by Anderson's vertical-stress model. Oblique faulting is relatively uncommon in shallow crustal seismicity (less than 5% of well-constrained events shallower than 30 km), but it becomes more prevalent in deeper or tectonically complex settings where stress rotations occur. These faults are commonly observed in oblique tectonic environments, such as subduction zones with non-perpendicular convergence or pull-apart basins within strike-slip systems. For instance, the in New Zealand serves as a continental transform boundary with significant oblique slip, accommodating both dextral strike-slip and reverse dip-slip components along a ~50° inclined plane, driven by the oblique relative motion between the Pacific and Australian plates. Similarly, in the , an oblique rift system, faulting includes hybrid structures along basin margins, where northwest-southeast extension combines with strike-slip to form moderate-dipping normal-oblique faults, facilitating the rapid propagation of the rift. These examples illustrate how oblique faults derive from Andersonian principles but adapt to inclined stress fields in real-world tectonics.

Applications in Geology

Tectonic regimes and stress indicators

Anderson's theory of faulting provides a foundational framework for interpreting tectonic regimes by linking the orientation of principal stresses to dominant fault types at or near the Earth's surface, where one principal stress is assumed to be vertical due to gravitational loading. In the extensional regime, characteristic of divergent plate boundaries like mid-ocean ridges and rift zones, the maximum principal stress (σ₁) is vertical, promoting normal faults that dip at approximately 60° and strike perpendicular to the minimum principal stress (σ₃). Compressional regimes, prevalent at convergent plate margins such as subduction zones, feature a vertical σ₃, resulting in low-angle reverse and thrust faults dipping at around 30° toward the direction of σ₁. The transcurrent or strike-slip regime, associated with transform boundaries like the , has a vertical intermediate principal stress (σ₂), yielding near-vertical faults with horizontal slip parallel to σ₁ or σ₃. These associations enable geologists to infer plate boundary dynamics from observed fault patterns, as validated by global seismic catalogs showing dominance of these modes in shallow crust (0–30 km depth). Stress indicators derived from Anderson's theory allow reconstruction of both contemporary and ancient stress fields. Fault orientations serve as primary proxies, with slip directions aligning to maximize shear stress on optimally oriented planes, typically at 30° to σ₁ under Coulomb failure criteria. Earthquake focal mechanisms provide dynamic indicators, where the P-axis (axis of maximum compression) approximates σ₁ and the T-axis (maximum extension) aligns with σ₃; for instance, vertical P-axes in normal faulting events confirm extensional regimes, while vertical T-axes indicate compression. These axes, derived from moment tensor inversions of seismic data, reveal global stress patterns consistent with plate tectonics, such as σ₁ orientations parallel to convergence directions in orogenic belts. Paleostress analysis employs multiple fault sets to uncover stress evolution, particularly through Andersonian conjugate pairs—symmetrical faults intersecting at ~60° with slip senses revealing the bisected acute angle as σ₁ direction. Graphical methods, like the dihedral angle approach, superimpose compression and extension quadrants from fault-slip data on stereonets to constrain principal axes, assuming contemporaneous deformation and no significant rotation. Such analyses of conjugate sets in sedimentary basins, for example, have documented regime shifts from extension to compression during orogenic cycles, providing timelines of tectonic history. The World Stress Map (WSM) project exemplifies global application, compiling 100,842 indicators—including fault-slip data and focal mechanisms—to map horizontal stress orientations (SHmax) and regimes worldwide as of 2025. By integrating WSM data with GPS-derived plate motion models, studies show that ~70% of continental stress aligns with absolute plate velocities, reinforcing Anderson's predictions; for instance, extensional regimes dominate in the Basin and Range Province, corroborated by both fault patterns and seismic P-T axes. This synthesis highlights how Andersonian theory bridges geological observations with geodetic data for plate-scale stress modeling, with recent database expansions enhancing resolution and coverage. Quantitative inversion methods compute principal stresses from fault data under Andersonian assumptions, reducing the stress tensor to three parameters (two directions and a shape ratio). Techniques like Reches' (1987) least-squares inversion minimize misfit between observed slip vectors and those predicted by the stress tensor on fault planes, yielding σ₁/σ₃ orientations and ratios (e.g., R = (σ₂ - σ₃)/(σ₁ - σ₃)). Simpler graphical inversions on stereonets transform data to identify great-circle distributions normal to the reduced stress vector, enabling separation of heterogeneous datasets into regime-specific subsets, as applied to Alpine fault arrays. These methods assume frictional sliding (μ ≈ 0.6) and vertical principal stress, providing robust estimates for tectonic interpretation.

Seismicity and earthquake mechanisms

Anderson's theory of faulting provides a framework for understanding fault reactivation in seismicity, where pre-existing faults oriented according to the theory—typically at angles near 30° to 45° to the principal stress directions—fail when the resolved shear stress approaches the , often requiring only a small increase in differential stress to trigger earthquakes. This reactivation is particularly relevant in the brittle upper crust, where frictional sliding on favorably oriented planes dominates seismic slip, as supported by analyses of borehole stress measurements and global earthquake catalogs. Focal mechanisms, visualized as "beach ball" diagrams, illustrate the orientations of principal stresses (σ₁ and σ₃) and slip directions consistent with Anderson's classifications, with the compressional (P) and tensional (T) axes aligning to show fault types. For instance, in extensional rifts, normal faulting mechanisms display a subvertical σ₁ (P axis near vertical) and horizontal σ₃, producing beach balls with white quadrants indicating extension on near-45° dipping planes, as observed in mid-oceanic ridge seismicity. Similarly, strike-slip events show horizontal σ₁ and σ₃ with vertical σ₂, yielding beach balls with alternating black and white quadrants on vertical nodal planes, while reverse mechanisms in convergent settings feature subhorizontal σ₁ and vertical σ₃. These patterns confirm the theory's prediction that one principal stress is vertical in shallow seismicity (<30 km depth), with dip-slip and strike-slip events comprising 32%–70% of well-constrained global data. Prominent examples demonstrate the theory's application to major earthquakes. The 1906 San Francisco earthquake (M_w 7.9) exemplified strike-slip faulting along the , with right-lateral rupture over 430 km and average slip of ~4 m, aligning with Anderson's transcurrent regime where vertical stress is intermediate and horizontal maximum stress is subparallel to the fault trace. In contrast, the 1989 Loma Prieta earthquake (M_w 6.9) involved reverse-oblique slip on a steep (~70°) northeast-dipping plane within a restraining bend of the , representing a non-ideal Andersonian orientation that required higher differential stress or lower friction for failure compared to optimal ~30°–45° dips in pure compression. The theory informs seismic hazard assessment by enabling predictions of slip directions from inferred stress orientations and aftershock patterns through Coulomb stress modeling. In stress shadows—regions of decreased Coulomb failure stress following a mainshock—aftershocks are suppressed on nearby Andersonian faults, as seen in reduced seismicity rates post-1906 along parts of the San Andreas, allowing probabilistic forecasts of triggered versus inhibited events. Centennial analyses in 2008 validated these concepts using the Global Centroid Moment Tensor catalog (1976–2004), confirming that shallow crustal earthquakes predominantly follow Andersonian faulting, with rare oblique mechanisms (<10% of events) and nodal plane dips clustering around theoretical values (e.g., ~45° for normal faults), thus underscoring the theory's robustness for interpreting global seismicity patterns.

Limitations and Extensions

Violations of assumptions

Anderson's theory of faulting assumes that one of the principal stresses is vertical, arising from the free surface boundary condition at the Earth's surface, which leads to predictions of fault dips around 60° for normal faults and 30° for reverse faults relative to the principal stress axes. This assumption of vertical principal stresses is violated in various geological settings, such as near major faults, salt domes, and subduction zone interfaces, where the maximum principal stress (σ1) can become tilted due to local perturbations from structural features or density contrasts. For instance, around salt domes, rising salt bodies induce radial and circumferential stress fields that rotate principal stress directions away from vertical, leading to non-Andersonian fault orientations in the overlying sediments. In subduction zones, the underthrusting of oceanic plates causes σ1 to tilt subhorizontally toward the trench, deviating from the vertical orientation expected in Anderson's model. Observational evidence for these tilted stresses comes from borehole breakouts and hydraulic fractures; in vertical boreholes, breakouts form perpendicular to the maximum horizontal stress, but when principal stresses are inclined, breakout orientations vary with depth and indicate non-vertical σ1, as documented in profiles near the San Andreas Fault where stress directions fluctuate and tilt. Similarly, hydraulic fractures propagate perpendicular to the minimum principal stress, revealing tilted axes in proximity to fault zones. Another key violation involves low-angle normal faults (LANFs), which dip at less than 30° and contradict the theory's prediction of ~60° dips for normal faults under vertical σ1 and horizontal extension. These structures are prominent in metamorphic core complexes, where large-magnitude extension exhumes deep crustal rocks along shallowly dipping detachments. A representative example is the Snake Range décollement in east-central Nevada, a Tertiary low-angle normal fault (~20-25° dip) that accommodated over 36 km of eastward displacement, separating ductilely deformed mid-crustal rocks from brittle supracrustal layers, as evidenced by geological mapping and geophysical constraints. Such LANFs pose a mechanical paradox because frictional sliding on low-angle planes under Andersonian stress states would require unrealistically high differential stresses or strength drops. Pore fluid pressure further violates the theory's assumptions by reducing effective normal stress on fault planes, enabling slip on low-angle thrusts that dip below the predicted 30°. Elevated pore pressures, often approaching lithostatic levels in geosynclinal sediments, lower the effective stress coefficient (λ = pore pressure / overburden pressure) toward 1, allowing overthrust sheets to move despite shallow dips, as originally proposed for large-scale thrusting. This mechanism explains anomalies like low-angle thrusts in fold-and-thrust belts where Anderson's dry-rock friction model predicts instability. In the system, deviations from Andersonian stress orientations, including rotated principal axes and suboptimal fault angles, are partly attributed to such fluid effects combined with local stress perturbations, as inferred from earthquake focal mechanisms and in-situ measurements showing horizontal compression at acute angles to the fault trace.

Modern refinements and alternatives

Subsequent developments have refined Anderson's theory by incorporating more realistic frictional behaviors observed in laboratory experiments. Byerlee's law provides an empirical description of rock friction on fault surfaces, stating that the shear stress required for sliding is approximately 0.85 times the normal stress for stresses above 200 MPa, and about 50 MPa for lower stresses, largely independent of rock type. This law extends the Coulomb friction assumption in Anderson's model by specifying friction coefficients (typically 0.6–0.85) that better match natural fault strengths in the upper crust. Further refinements account for the role of pore fluid pressure in reducing effective stress on faults, following Terzaghi's principle. The effective normal stress is given by \sigma' = \sigma - P, where \sigma is the total normal stress and P is the pore pressure, which lowers the frictional resistance and can trigger slip at lower differential stresses than predicted by Anderson's hydrostatic assumptions. Elevated pore pressures, often due to fluid migration or tectonic loading, explain seismicity in otherwise stable regimes and are integrated into modern fault stability assessments. Three-dimensional extensions of Anderson's theory relax the assumption of vertical principal stresses, allowing for tilted stress orientations that arise from complex tectonic interactions. Numerical models, such as those implemented in the FLAC (Fast Lagrangian Analysis of Continua) software, simulate fault development in heterogeneous media by incorporating full 3D stress tensors and material nonlinearity, revealing how inclined principals lead to non-conjugate fault patterns. These models demonstrate that deviations from Andersonian geometries occur in regions with pre-existing weaknesses or lateral variations in crustal properties. Alternative approaches hybridize Anderson's conjugate fault predictions with the to handle intermediate stress effects and curved failure envelopes. In these models, fault orientations are determined by the full , accommodating hybrid fractures that combine extension and shear modes when the stress path intersects non-linear envelopes. Elastic dislocation theory provides another alternative for modeling fault evolution, treating faults as dislocations in an elastic medium to simulate progressive growth through linkage and propagation. This framework captures displacement accumulation and segment interaction, explaining irregular fault shapes beyond simple planar assumptions. Integrations with plate tectonics have mapped global stress provinces, where Andersonian regimes align with boundary forces like ridge push and slab pull. Zoback's analysis delineates first-order stress patterns over scales >500 km, showing compressional provinces in continental interiors consistent with plate-scale driving mechanisms. Recent advancements employ for stress inversion from fault-slip data, using neural networks to classify polyphase deformation and infer paleostress orientations with higher accuracy than traditional methods. As of 2025, Anderson's theory remains foundational for interpreting fault but is augmented by geodetic observations from GPS and InSAR, which provide real-time validation of stress orientations and slip vectors. These techniques reveal subtle deviations from ideal Andersonian dips, such as in global fault datasets, while confirming the theory's predictions in many active margins.

References

  1. [1]
    The dynamics of faulting - Lyell Collection
    THE DYNAMICS OF FAULTING. 387. XLIL. The Dynamics of Faulting. B y ERNEST M . ANDERSON,. M.A., B.Sc, H.M. Geological Survey. (Read 15th March 1905.) IT has ...
  2. [2]
  3. [3]
    Dynamics of Faulting and Folding - SpringerLink
    Anderson, E. M.: The Dynamics of Faulting and Dyke Formation with Applications to Britain. Edinburgh: Oliver & Boyd Ltd. 1942. Google Scholar. Hafner, W.: Bull.
  4. [4]
    Stress, faulting, fracturing and seismicity: the legacy of Ernest ...
    Aug 9, 2012 · Field geologists commonly find that low-displacement faults fit into one or the other of Anderson's three classes, and that conjugate brittle ...Missing: classification | Show results with:classification
  5. [5]
    Catalog Record: The dynamics of faulting and dyke formation...
    The dynamics of faulting and dyke formation with applications to Britain. Author: Anderson, E. M. (Ernest Masson), 1877-1960. Published: 1951. The dynamics of ...
  6. [6]
    The Dynamics of Faulting and Dyke Formation with Applications to ...
    Title, The Dynamics of Faulting and Dyke Formation with Applications to Britain ; Author, Ernest Masson Anderson ; Edition, 2 ; Publisher, Oliver and Boyd, 1951.
  7. [7]
    [PDF] The Mechanics of Earthquakes and Faulting
    3.1.1 Anderson's theory of faulting 101. 3.1.2 Hubbert–Rubey theory of overthrust faulting 104. 3.1.3 Stress in the crust, fault reactivation, and friction ...
  8. [8]
    Andersonian and Mohr-Coulomb Theory of Faulting - Rock Fracture
    The Andersonian theory of faulting (Anderson 1951) is probably the earliest theory regarding the mechanism of faulting. It is based on the Coulomb failure ...
  9. [9]
    Stress and Strain - SERC (Carleton)
    Oct 10, 2007 · Stress is a force acting on a rock per unit area. · Strain is a change in shape or size resulting from applied forces (deformation). · Structures ...
  10. [10]
    Stress and Strain - Rock Deformation
    Rock responds to stress differently depending on the pressure and temperature (depth in Earth) and mineralogic composition of the rock.
  11. [11]
    Stress, Strain and Structures – Geology 101 for Lehman College ...
    Stress is the force applied to an object. In geology, stress is the force per unit area that is placed on a rock. Four types of stresses act on materials.Missing: fundamentals | Show results with:fundamentals
  12. [12]
    9.1 Stress and Strain – Physical Geology: An Arizona Perspective
    Rock can respond to stress in three ways: it can deform elastically, it can deform plastically, and it can break or fracture.Missing: fundamentals | Show results with:fundamentals
  13. [13]
    Crustal Deformation and Earthquakes – Introduction to Earth Science
    Strain in rocks can be represented as a change in rock volume and/or rock shape, as well as fracturing the rock. There are three types of stress: tensional, ...
  14. [14]
    The Brittle‐Plastic Transition, Earthquakes, Temperatures, and ...
    Apr 30, 2020 · Earthquakes in tectonically active regions can occur in regions where temperatures reach 700–900°C Strain rates versus highest temperatures ...
  15. [15]
    [PDF] Rock mass properties - Rocscience
    Many rocks show a significant strength decrease with increasing moisture content and tests on samples, which have been left to dry in a core shed for several.
  16. [16]
    [PDF] Calculating the cohesion and internal friction angle of volcanic rocks ...
    Jan 30, 2022 · Cohesion and friction angle are the two defining material variables for the Mohr-Coulomb failure criterion. Although these can be determined ...
  17. [17]
    [PDF] Stress & Strain: A review - UBC EOAS
    The normal and shear stress components are the normal and shear forces per ... normal stress (∆N/∆A) and the shear stress (∆S/∆A) as the area ∆A ...
  18. [18]
    Stress - Course Listing | Earth and Atmospheric Sciences
    Lithostatic stress is similar, but due to the density of overlying rock, not water. It's an appropriate description for parts of the crust that have remained ...
  19. [19]
    Seeking Anderson's faulting in seismicity: A centennial celebration
    Oct 1, 2008 · Anderson proposed the simplest, yet realistic, theory that relates faulting to the state of stress to the Edinburgh Geological Society about a ...
  20. [20]
    Tensors: Stress, Strain and Elasticity - SERC (Carleton)
    Jan 19, 2012 · The Eigen vectors are the principle stress directions known as the maximum, intermediate and minimum principle stresses respectively; in geology ...
  21. [21]
    Mohr circles and stress transformation - Geological Digressions
    Mar 23, 2024 · Mohr circles are a simple graphical system for calculating normal stress and shear stress in rotating (deforming) structural elements.
  22. [22]
    [PDF] Notes VII Faults 1
    Rock mechanics and Anderson's theory of faulting give us a first order picture of how the types and orientations of faults are related to the orientations ...
  23. [23]
    Mohr-Coulomb failure criteria - Geological Digressions
    Apr 3, 2024 · On the Mohr circle this is plotted as 2θ. In a conjugate fracture set there will be two failure envelopes – one at or close to the maximum ...
  24. [24]
    [PDF] The dynamics of faulting - Semantic Scholar
    The dynamics of faulting · E. M. Anderson · Published in Transactions of the Edinburgh… · Geology.
  25. [25]
    [PDF] Mechanics of faulting
    Anderson's theory of faulting. ➢ Condition: near the free surface one of the principal stresses is vertical,. ➢ Three major classes of faults (reverse, normal, ...
  26. [26]
    Mohr–Coulomb Failure Criterion | Rock Mechanics and Rock ...
    Jul 3, 2012 · ... principal stress plane or the Mohr diagram, which is often convenient. ... Jaeger JC, Cook NGW (1979) Fundamentals of Rock Mechanics, 3rd edn.<|control11|><|separator|>
  27. [27]
    L. Tectonic Environments of Faulting – Geological Structures
    Anderson's (1905) theory of faulting starts from some basic facts about stress, and leads to a classification of tectonic environments into fault regimes.
  28. [28]
    Seeking Anderson's faulting in seismicity: A centennial celebration
    Oct 1, 2008 · First, it predicts slip to be along the fault plane strike or dip direction only and therefore does not account for frequently observed oblique ...
  29. [29]
    [PDF] Faults and Faulting
    Anderson, E. M., 1951. Dynamics of faulting and dyke formation. Oliver and Boyd: Edinburgh. Bonnet, E., Bour, O., Odling, N. E., Davy, P., Main, I.,. Cowie ...
  30. [30]
  31. [31]
    [PDF] Introduction San Andreas Fault: An Overview
    Stream alignment may be a result of strike-slip fault motion or from the erosion of sheared and pulverized rock along the fault trace. Bay Area examples include ...
  32. [32]
    The Mechanics of Oblique Slip Faulting | Geological Magazine
    May 1, 2009 · The various mechanisms which could cause oblique slip faulting are briefly reviewed. It is thought that such faulting may frequently arise ...
  33. [33]
    New insights from LiDAR, central Alpine fault, New Zealand
    Mar 9, 2017 · At the first order (<106–104 m), the Alpine fault is a remarkably straight and unpartitioned structure controlled by inherited and active ...
  34. [34]
    Deformation produced by oblique rifting - ScienceDirect
    Our results explain the presence and trends of oblique-slip and strike-slip faults along the margins of the Gulf of California and the oblique trend (relative ...
  35. [35]
    [PDF] PALEO-“STRESS” ANALYSIS FROM FAULT DATA
    Transtension reports the simultaneous occurrence of strike-slip faults, normal faults and oblique-slip normal faults. Tectonic regimes near the transition ...<|separator|>
  36. [36]
    World Stress Map
    The World Stress Map (WSM) is a global compilation of information on the crustal present-day stress field maintained since 2009.Download · Data · Casmo · PublicationsMissing: Anderson | Show results with:Anderson
  37. [37]
    [PDF] DETERMINATION OF THE TECTONIC STRESS TENSOR FROM ...
    Anderson, E. M., The Dynamics of Faulting, 183 pp., Oliver and Boyd, London, 1942. Angelier, J., Determination of the mean principal directions of stresses ...Missing: summary | Show results with:summary
  38. [38]
    A simple stress inversion of fault/slip data assuming Andersonian ...
    Apr 24, 2004 · The parameter (or sigma) space may be reduced from five to three dimensions if an Andersonian stress state (having a vertical principal ...
  39. [39]
    [PDF] Earthquakes on Compressional Inversion Structures –
    Sep 30, 2016 · Non-Andersonian Steep Reverse Faults (dips > 45°). Page 11. • Mw > 5.5 ... 1989 M6.9 Loma Prieta EQ.<|separator|>
  40. [40]
    Stress Triggers, Stress Shadows, and Implications for Seismic Hazard
    the slip direction of the second earthquake;/X o'n is the change in normal stress due to the first earthquake, resolved in the direction orthogonal to the ...
  41. [41]
    How do stress perturbations near salt bodies induce difficulty in salt ...
    May 22, 2019 · This study suggests that the near-salt stress perturbations should not be ignored in salt imaging, and that geomechanical stress modelling is ...
  42. [42]
    Stress orientation profile to 3.5 km depth near the San Andreas Fault ...
    Apr 10, 1992 · The variations in breakout orientation at depth seem to reflect stress fluctuations associated with active faults penetrated by the drill hole.<|separator|>
  43. [43]
    Stress orientations in northern and central California: Evidence for ...
    Mar 29, 2003 · We determined orientations of principal stresses around the San Andreas fault (SAF) system in the greater San Francisco Bay Area and regions ...Missing: deviations | Show results with:deviations
  44. [44]
    [PDF] The mechanical paradox of low-angle normal faults
    Anderson–Byerlee frictional fault reactivation theory, i.e. an extending crust characterised by vertical trajectories of σ1 and normal faults possessing ...
  45. [45]
    The Snake Range Décollement interpreted as a major extensional ...
    Geological and geophysical constraints suggest that the Snake Range décollement of east-central Nevada is a major Tertiary low-angle normal fault zone.
  46. [46]
    Role of fluid pressure in mechanics of overthrust faulting - USGS.gov
    The hypothesis that large-scale overthrusting is facilitated by abnormal fluid pressures which, in turn, are associated with geosynclinal deposition is applied ...
  47. [47]
    New Evidence on the State of Stress of the San Andreas Fault System
    Contemporary in situ tectonic stress indicators along the San Andreas fault system in central California show northeast-directed horizontal compression that ...
  48. [48]
    Friction of rocks | Pure and Applied Geophysics
    At low stress rock friction is strongly dependent on surface roughness. At high normal stress that effect is diminished and the friction is nearly independent ...Missing: law | Show results with:law
  49. [49]
    Influence of Effective Stress and Pore Fluid Pressure on Fault ...
    Oct 15, 2020 · The data show a dependence of shear stress and—apparently—friction coefficient on Terzaghi stress, that is well addressed by the Coulomb law.
  50. [50]
    Modelling fault reactivation and velocity dependent friction with ...
    Oct 24, 2015 · In this paper, the capabilities of TOUGH has been extended with the sequential coupling to the geomechanical finite difference FLAC solver.
  51. [51]
    (PDF) On the importance of 3D stress state in 2D earthquake rupture ...
    To this day, Anderson's theory is the only framework that can be used to link fault generation and reactivation to the three-dimensional stress field. In this ...
  52. [52]
    Use of Mohr diagrams to predict fracturing in rock masses, with ...
    Jun 14, 2024 · Hybrid fractures develop by synchronous extension and shear modes, and occur if the Mohr circle touches the curved part of the failure envelope ...
  53. [53]
    Elastic Dislocation Modeling Helps Identify Subseismic Natural ...
    Feb 1, 2025 · The main goal of this research work was to determine the subseismic faults and fracture corridors and their characteristics, including density ...
  54. [54]
    [PDF] Elastic dislocation modeling of wrinkle ridges on Mars
    Multiple-fault kinematic models have been intro- duced that involve deformation over a blind or surface- breaking thrust fault and the development of a ...
  55. [55]
    First‐ and second‐order patterns of stress in the lithosphere: The ...
    Jul 30, 1992 · The “first-order” midplate stress fields are believed to be largely the result of compressional forces applied at plate boundaries, primarily ridge push and ...
  56. [56]
    A Deep Learning‐Aided Workflow for Decoding the Stress Regime ...
    Mar 12, 2025 · The stress inversion results provide additional and complementary insight into earthquake processes in Nevada. We invert the entire dataset in ...
  57. [57]
    Global distribution of earthquake-fault dip angles - Oxford Academic
    Anderson's theory assumptions require that one of the three principal stress axes is nearly orthogonal to the Earth's surface, and faulting should occur at 45° ...
  58. [58]
    Scripps Institution of Oceanography expands geodetic program
    Nov 5, 2025 · Her ongoing work includes applying Andersonian faulting theory to assess whether fractures formed before or after earthquakes. Andersonian ...