Fact-checked by Grok 2 weeks ago

Shear force

Shear force is the internal force acting parallel to a cross-section of a structural member, such as a , that tends to cause one portion of the member to slide relative to another, arising from transverse loads applied along the member's length. It is defined as the algebraic sum of all vertical forces acting on one side of the section, ensuring vertical equilibrium when added to the section. In , shear forces are fundamental to analyzing and other flexural members, where they combine with bending moments to determine internal stresses and ensure under loads like , , or seismic forces. The magnitude of shear force varies along the length of a , decreasing or increasing based on distributed or concentrated loads; for instance, a uniform distributed load produces a linear variation in shear force, while a point load causes an abrupt change. Engineers use shear force diagrams (SFDs) to visualize this variation, plotting shear force against position along the to identify maximum values that could lead to shear failure if not properly resisted by the material's . The relationship between shear force and bending moment is governed by differential equations derived from equilibrium: the rate of change of shear force equals the negative of the load intensity (dV/dx = -q), and the rate of change of bending moment equals the shear force (dM/dx = V). This interplay is critical in design, as excessive shear can cause diagonal cracking or failure in concrete beams or yielding in steel, necessitating reinforcements like stirrups or web stiffeners. Overall, understanding and calculating shear forces enables safe and efficient structural systems across civil, mechanical, and applications.

Fundamentals

Definition

Shear force is defined as the internal transverse force component within a structural member, such as a , that acts parallel to the cross-sectional area and tends to cause one portion of the member to slide relative to an adjacent portion, resulting in shearing deformation. This force arises from external loads applied perpendicular to the member's longitudinal and represents the resultant of the distributed internal shearing forces across the section. In contrast to axial forces, which act normal to the cross-section along the length of the member and produce direct or , shear force operates tangentially to induce parallel between layers. forces, manifested as moments, cause rotational deformation leading to , whereas shear force specifically promotes a sliding or shearing action without primary . As a quantity, shear force is directed to the of the structural member but lies within the of the cross-, with its and determined by the of transverse components on either side of the . A representative example occurs in a cantilever fixed at one end and subjected to a transverse point load at the free end; in this case, the shear force remains constant along the beam's length, equal in to the applied load but opposite in to counteract the shearing tendency.

Historical Development

The concept of shear force, as a tangential force causing sliding or deformation within materials, received early recognition through Leonardo da Vinci's investigations into and material failure during the late 15th and early 16th centuries. Da Vinci conducted systematic experiments on sliding between surfaces, observing how tangential forces led to resistance and eventual failure in materials like wood and metal, laying groundwork for understanding shear-related phenomena in . In the , advanced the study of in practical contexts through his 1773 essay on the application of maxima and minima to problems in . Coulomb analyzed and along sliding planes in soils and , deriving equations for earth pressure and resistance that formed the basis of and influenced later structural analyses. Meanwhile, the Euler-Bernoulli beam theory, developed in the 1740s by Leonhard Euler and , provided a foundational framework for bending but initially overlooked deformation, assuming plane sections remained perpendicular to the under load. The 19th century saw formal incorporation of effects into beam and prism theories. Claude-Louis Navier's 1826 work on elastic beams introduced more comprehensive elasticity principles, though was still approximated. Adhémar Jean Claude Barré de Saint-Venant extended this in the , notably through his 1856 memoir on the torsion of prisms, where he rigorously derived distributions and warping effects in non-circular cross-sections, bridging torsion and flexural . Jacques Antoine Charles further refined the approach in 1859 by explicitly including flexibility and rotary inertia in beam equations, correcting the limitations of Euler-Bernoulli for shorter beams. Concurrently, James Clerk Maxwell's developments in graphical (1864–1870) enabled the first explicit visualizations of internal forces, including diagrams, through figures that graphically represented in frames and beams. By the early , shear deformation gained prominence in . Stephen Timoshenko's seminal papers in 1921 and 1922 introduced a corrected accounting for both shear deformation and rotary , using a shear correction factor to match three-dimensional elasticity solutions; this work, published in the and elsewhere, became integral to modern textbooks and analyses of thick beams.

Mechanics of Shear Force

Shear Force in Beams

In beam theory, shear force arises from the transverse loading on a and is determined through static considerations. For a subjected to a distributed transverse load w(x) (positive downward), the shear force V(x) at any satisfies the derived from vertical on an infinitesimal element: \frac{dV}{dx} = -w(x). This relation indicates that the rate of change of shear force along the length equals the negative of the distributed load intensity, as the load opposes the shear's variation. Integrating this yields V(x) = V_0 - \int_{x_0}^x w(\xi) \, d\xi, where V_0 is the shear at reference point x_0. The shear force is also directly related to the M(x) through moment equilibrium. Consider an infinitesimal beam element of length dx with shear forces V and V + dV acting on its ends, and s M and M + dM. Summing moments about one end (neglecting higher-order terms like dV \cdot dx) gives dM = V \, dx, leading to the \frac{dM}{dx} = V. This shows that the slope of the diagram equals the shear force at any point, with providing M(x) = M_0 + \int_{x_0}^x V(\xi) \, d\xi, where M_0 is the at x_0. These equations hold in regions of continuous loading but exhibit discontinuities at concentrated forces or moments. A consistent sign convention is essential for applying these relations. Positive shear force is defined such that it causes rotation of the segment to the left of the section cut, or equivalently, tends to displace the left portion upward relative to the right. This aligns with deformations where the beam cross-section rotates under positive . Bending moments follow a complementary , with positive moments producing compression on the top fibers. Boundary conditions for shear force depend on the support type. At a free end, where no transverse restraint exists, the shear force is (V = 0), as no external force balances any internal shear. At a fixed or pinned , the shear force equals the vertical reaction force provided by the , determined from overall of the . For cantilever beams fixed at one end and free at the other, shear is at the free end and equals the total transverse load at the fixed end. A representative example is a simply supported of L under a uniform distributed load w (per unit ). The reactions at each are \frac{wL}{2}, so the shear force is maximum at the supports, with V = \frac{wL}{2} (positive at the left support and negative at the right, per ). The shear varies linearly from +\frac{wL}{2} at x = 0 to -\frac{wL}{2} at x = L, crossing zero at midspan where V(x) = \frac{wL}{2} - w x = 0 (i.e., x = \frac{L}{2}). This distribution highlights how shear peaks near supports in such configurations.

Shear Force Diagrams

Shear force diagrams provide a graphical representation of how the internal shear force varies along the length of a structural member, such as a , enabling engineers to visualize force distribution for and . These diagrams plot the shear force V against the position x along the member, typically drawn below a free-body of the for reference. To construct a shear force diagram, support reactions are first determined using static equilibrium equations, such as \sum F_y = 0 and \sum M = 0. The diagram begins at one end of the beam, often the left , where V equals the vertical reaction force (positive if upward on the left face). As position x progresses, the shear force is updated by considering loads to the left of each section: for a concentrated point load, V experiences a sudden vertical discontinuity, jumping upward by the load magnitude if the load is upward or downward if downward; for distributed loads, V changes with a slope equal to the negative of the load intensity w(x), resulting in linear segments for uniform distributed loads (constant slope -w) and curved segments for varying loads, such as parabolic curves for linearly varying triangular loads. The process continues section by section until the opposite end, where V should return to zero or match the end reaction for . Key features of shear force diagrams include abrupt discontinuities at points of concentrated loads, reflecting instantaneous force changes, and smooth transitions with defined slopes under distributed loading, where uniform loads produce straight linear variations and more complex loads yield or higher-order curves. These characteristics highlight regions of constant shear (horizontal lines between loads) and accelerating changes due to distributed effects. Shear force diagrams are commonly integrated with bending moment diagrams, plotted on the same horizontal axis but with vertical scales adjusted for each; the area beneath the shear curve between any two points equals the change in \Delta M over that interval, providing a direct link for comprehensive internal force analysis. This complementary visualization aids in verifying calculations and understanding load transfer. In practice, shear force diagrams are essential for identifying locations of maximum shear force, which informs the design of cross-sections to withstand shear stresses and prevents ; for instance, maximum |V| often occurs near , guiding placement in structural elements. A representative example is a simply supported of length L subjected to a uniform distributed load with total magnitude P = wL. The reactions are each P/2 upward. The shear force starts at +P/2 just inside the left , decreases linearly with slope -w to zero at the beam's center (x = L/2), and then continues linearly to -P/2 just inside the right , forming a symmetric triangular variation. This diagram reveals the maximum shear of P/2 at the , critical for design checks.

Shear Stress and Strength

Relation to Shear Stress

Shear force in a structural member induces across its cross-section, which represents the internal resistance to the applied transverse loading. The simplest approximation for is the average value, given by \tau_{avg} = V / A, where V is the shear force and A is the cross-sectional area to . This average is useful for preliminary calculations but overlooks the non-uniform distribution of , which varies significantly with position in the cross-section, particularly in beams under . For a more precise analysis, especially in beams, the shear stress distribution is determined using Jourawski's formula, derived by Dmitrii Ivanovich Zhuravskii in 1855: \tau = \frac{V Q}{I t}, where Q is the first moment of the area above (or below) the point of interest about the , I is the second moment of area () of the entire cross-section, and t is the width (thickness) at the location where \tau is calculated. This formula arises from considerations of horizontal forces in beam elements and provides the longitudinal shear stress, which equals the transverse by complementarity. The derivation and application of Jourawski's formula rely on key assumptions from beam theory, including a linear elastic, isotropic material behavior, small deformations, and the plane sections remaining plane after deformation (as per the Bernoulli-Euler hypothesis). These ensure that normal stresses due to do not interfere with the shear distribution calculation. In a rectangular cross-section, the shear stress distribution is parabolic, with the maximum \tau_{max} = \frac{3V}{2A} occurring at the neutral axis and zero at the top and bottom fibers. For thin-walled sections, such as those in closed tubular members, the shear stress is approximately uniform across the wall thickness, simplifying design assessments. Consider an subjected to vertical shear force V; the stress profile shows low values in the flanges (near zero at outer edges) and peaks in the web, where Q is largest due to the greater area contribution away from the . This distribution highlights why the web carries most of the shear load in such efficient structural shapes.

Shear Failure Criteria

Shear failure criteria define the conditions under which a or fails due to excessive , typically when the applied exceeds the material's capacity to resist deformation or . These criteria are essential for predicting ing or ultimate in engineering design, particularly for ductile and brittle materials under complex loading. The Tresca yield criterion, also known as the maximum theory, posits that yielding occurs when the maximum shear stress (\tau_{\max}) in the reaches half the strength (\sigma_y) obtained from a uniaxial test, expressed as \tau_{\max} = \frac{\sigma_y}{2}. This criterion is conservative and particularly suitable for materials where shear is the dominant mode, such as in ductile metals under loading. In contrast, the , or distortion energy , predicts failure based on the accumulation of distortional , where yielding initiates when the effective equates to the distortional energy at in uniaxial . For conditions, this yields a critical of \tau = \frac{\sigma_y}{\sqrt{3}} \approx 0.577 \sigma_y, providing a less conservative estimate than Tresca for most ductile materials. This better aligns with experimental data for multiaxial states in metals. For ultimate shear strength, ductile materials like steels typically exhibit a shear capacity of approximately 0.75 \sigma_{uts} based on empirical relations for design purposes. For brittle materials like cast irons, the ultimate shear strength often exceeds the ultimate tensile strength (e.g., approximately 1.5 \sigma_{uts} for gray cast iron), though with sudden fracture and no ductility; ceramics vary depending on type. Several factors influence , including material type, where steels achieve about 0.75 times their tensile strength in shear, while polymers or composites vary significantly based on . Temperature reduces by enhancing atomic mobility and softening the material lattice, with linear decreases observed in metals up to elevated levels. also affects capacity, as higher rates increase in rate-sensitive materials like metals through dislocation dynamics, though excessive rates can induce . In , shear failure manifests as diagonal cracks due to principal tensile stresses at 45 degrees to the plane, prompting the use of stirrups or shear reinforcement designed according to ACI 318 codes to enhance capacity by crossing potential crack paths. These reinforcements limit crack widths and transfer forces, ensuring ductile behavior over brittle failure. A representative example is punching shear failure in slabs, where concentrated loads from columns cause a conical failure surface; the critical perimeter for is typically at a distance of d/2 from the column face, where d is the effective depth, allowing of the against material limits to prevent localized rupture.

Applications and Examples

In

In , shear force plays a pivotal role in ensuring the and of civil structures such as bridges, buildings, and retaining walls, where it must be accurately assessed and mitigated to prevent catastrophic failures. codes provide standardized provisions for calculating and limiting allowable shear forces based on material properties and loading conditions. For instance, the AASHTO LRFD Bridge Design Specifications outline shear requirements for bridge components, including prestressed and non-prestressed girders, emphasizing the use of factored shear forces to determine capacities. Similarly, Eurocode 2 (EN 1992-1-1) governs shear in structures, specifying methods for verifying shear in beams and slabs without or with shear reinforcement, particularly for members subjected to and transverse forces. These codes incorporate load factors to account for uncertainties, such as the combination 1.2D + 1.6L for dead (D) and live (L) loads in strength , as per ASCE 7, ensuring that the shear force Vu does not exceed the nominal capacity φVn. To enhance shear resistance, engineers employ targeted reinforcement strategies tailored to the structural material and configuration. In reinforced concrete beams, vertical stirrups—typically made of deformed bars or welded wire—act as shear reinforcement by crossing potential diagonal tension cracks and tying the compression zone to the tension , thereby increasing the member's shear capacity beyond that of alone. Shear keys, often used in bridge abutments or precast connections, provide mechanical interlock to transfer horizontal shear forces across joints, preventing sliding under seismic or lateral loads. For steel plate girders, which are common in long-span bridges, transverse web stiffeners are installed at intervals along the web to prevent under high shear stresses, effectively subdividing the web into smaller panels that enhance post- shear strength and overall stability. Analysis methods for shear force vary by structure complexity, balancing computational efficiency with accuracy. Simplified , based on Euler-Bernoulli assumptions, is suitable for prismatic members like straight girders under uniform loading, allowing quick estimation of shear distribution via equilibrium equations. In contrast, finite element analysis (FEA) is essential for complex geometries, such as curved bridges or irregular frames, as it models three-dimensional states, including interactions with and torsion, providing more precise force predictions than classical methods. Modern structural design increasingly relies on software tools to automate shear force evaluations and ensure code compliance. Programs like SAP2000 facilitate integrated modeling, analysis, and design checks for shear in and elements, generating shear force diagrams, verifying reinforcement needs, and applying code-specific factors such as those from AASHTO or Eurocode 2 within a single environment. This approach not only streamlines workflows for large-scale projects but also allows iterative optimization to minimize material use while maintaining safety margins against shear-induced failure modes.

In Materials Testing

In materials testing, shear force resistance is evaluated through specialized experimental setups designed to isolate and quantify shear loading on specimens. Single shear tests involve applying a transverse load to a specimen fixed at one end and loaded at the other, creating a single plane, commonly used for thin wrought and cast aluminum alloy products to determine ultimate shear strengths per ASTM B831. Double shear tests, in contrast, load the specimen between two supports to produce two planes, distributing the force and reducing effects; this method is standardized for aluminum alloys under ASTM B769 to measure shear ultimate strengths more accurately for thicker materials. Torsion tests achieve a state of by twisting cylindrical or tubular specimens, eliminating normal stresses and providing fundamental data on shear behavior, particularly for ductile metals where radial gradients are analyzed. Shear force is typically measured using load cells integrated into testing machines, which convert the applied transverse or torsional load into electrical signals proportional to , while gauges bonded to the specimen surface capture deformation via changes in electrical . The ultimate is calculated as the maximum force V_{\max} divided by the effective shear area A, yielding \tau_u = V_{\max} / A, where this ratio establishes the material's capacity to withstand before failure. For polymers, the Iosipescu test employs a V-notched specimen loaded in to concentrate at the notch, minimizing interference and enabling accurate measurement of nonlinear response up to large deformations. In composites, the ±45° off-axis tensile test infers the in-plane by applying uniaxial tension to a laminate oriented at ±45° to the direction, where the dominates the response, as standardized in ASTM D3518 for polymer composites reinforced by high-modulus fibers. Data from these tests are interpreted through shear stress-strain curves, which plot against to distinguish the regime—characterized by linear recovery upon unloading—and the regime, where permanent deformation occurs, often showing hardening or necking in metals. These curves provide key parameters like the in the portion and the onset of yielding, aiding in for applications requiring specific deformation resistance. An example of assessing shear toughness under dynamic conditions is the instrumented variant for metals, where a notched specimen is struck by a , and load-time data from the instrumented allow estimation of fracture by analyzing the dominated by shear lip formation on the fracture surface.

References

  1. [1]
    [PDF] Shear Forces and Bending Moments in Beams
    Because shear forces and bending moments are the resultants of stresses distributed over the cross section, they are known as stress resultants and in ...Missing: definition | Show results with:definition
  2. [2]
    [PDF] Recitation #5 Understanding Shear Force and Bending Moment ...
    Shear force may be defined as "the algebraic sum of the loads to the left or right of a point (such that the addition of this force restores vertical equilbrium) ...
  3. [3]
    [PDF] 3. Shear Stress and Shear Strain
    a) Shear stress A shear force is applied to the end of a short, stubby member, where A is the cross- sectional area of the member. A mathematical cut is made ...
  4. [4]
    [PDF] mechanics of materials - Mercer University
    The resultant of the internal shear force distribution is defined as the shear of the section and is equal to the load P. • Corresponding internal forces act in ...
  5. [5]
    [PDF] Chapter 7 - NJIT
    What I mean by solving a structures problem for a beam is to compute the shear stress and bending moment within the beam given a set of external forces at ...Missing: definition | Show results with:definition
  6. [6]
    Understand Shear Forces [An Engineering Explanation]
    A shear force describes the internal force that acts parallel to a surface trying to slide or deform one part of a surface relative to another.What Is The Shear Force? · Shear Force Calculation · Shear Force Formulas
  7. [7]
    Comment on Leonardo da Vinci's Friction Experiments: An Old Story ...
    May 14, 2015 · Sliding inception is treated as a plastic yield failure mechanism, which allows static friction modeling under highly adhesive conditions.
  8. [8]
    [PDF] Coulomb, the artisan of modern geotechnical engineering - CFMS
    In his Essay (1773), Coulomb presented clear propositions on friction and cohesion on a sliding plane, giving birth to today's soil mechanics. Coulomb's theory ...
  9. [9]
    [PDF] Euler-Bernoulli Beams - Professor Terje Haukaas
    Aug 21, 2023 · The Euler-Bernoulli beam theory was established around 1750 with contributions from. Leonard Euler and Daniel Bernoulli.
  10. [10]
    [PDF] Who developed the so-called Timoshenko beam theory? - HAL
    Dec 20, 2019 · shear deformation effects were first introduced by Rankine [7] and rotary inertia effects by Bresse [10]. In his often-cited paper of 1921, ...<|control11|><|separator|>
  11. [11]
    On Torsion and Shear of Saint-Venant Beams - ResearchGate
    Aug 6, 2025 · Following Reddy (2007) and Wang et al. (2010), the shear correction factor of the local Timoshenko beam theory is adopted in the present paper.<|separator|>
  12. [12]
    None
    ### Summary of Shear Force and Bending Moment Derivation
  13. [13]
    Beam Stress & Deflection - MechaniCalc
    Sign Convention​​ The shear force at the section cut is considered positive if it causes clockwise rotation of the selected beam section, and it is considered ...Beam Analysis · Cross Section Properties · Beam Analysis Formula Sheet
  14. [14]
    Beams, Bending, and Boundary Conditions - The Geometry Center
    Apr 12, 1996 · w'''(L)=0 . This boundary condition models the assumption that there is no shearing force acting at the free end of the beam.
  15. [15]
    [PDF] DA6-BeamFormulas.pdf - Purdue Engineering
    R = reaction load at bearing point, lbs. V = shear force, lbs. W = total uniform load, lbs. w = load per unit length, lbs.
  16. [16]
    Shear and Moment Diagrams - Mechanics Map
    Shear and moment diagrams are graphical plots of internal forces and moments, used to find internal shear forces and bending moments in a beam.
  17. [17]
    Shear and Moment Diagrams – An Ultimate Guide - Engineering Skills
    Apr 7, 2025 · This tutorial provides a thorough introduction to shear forces, bending moments and how to draw shear and moment diagrams for beams and ...Building Shear and Moment... · Shear and Moment Diagram...
  18. [18]
    Shear Stress Equations and Applications - Engineers Edge
    This formula is also known as the Jourawski formula ... Beam shear is defined as the internal shear stress of a beam caused by the shear force applied to the beam ...Missing: derivation | Show results with:derivation
  19. [19]
    [PDF] Lecture 16 – Shear stress in beams
    Aug 20, 2025 · Equilibrium of beams​​ Determine the maximum shear stress in a pipe ( ). This equation can be used as an approximation for the maximum shear ...
  20. [20]
    Tresca Criterion - an overview | ScienceDirect Topics
    Tresca criterion states that a material point yields when the maximum shear stress at that point reaches the maximum shear stress in a uniaxial tension ...
  21. [21]
    11.8: Tresca Yield Condition - Engineering LibreTexts
    Mar 5, 2021 · In 1860 the French scientist and engineer Henri Tresca put up a hypothesis that yielding of the material occurs when the maximum shear stress ...
  22. [22]
    Von Mises Criterion ( Maximum Distortion Energy Criterion )
    This criterion is based on the determination of the distortion energy in a given material, ie, of the energy associated with changes in the shape in that ...
  23. [23]
    Maximum Shear Stress Criterion - 2023 - SOLIDWORKS Help
    For a condition of pure shear, von Mises stress criterion predicts failure at (0.577*yield strength), whereas the shear stress criterion predicts failure at 0.5 ...
  24. [24]
    Is there any relationship between shear strength and tensile strength ...
    Sep 19, 2013 · The relationship varies from 0.5 to 0.577, whether you use Tresca or von Mises criterion. Also, a value of 0.6 (which is an approximation of 0.577) is ...Missing: sigma_uts | Show results with:sigma_uts<|separator|>
  25. [25]
    Shear Strength - Eng-Tips
    Oct 6, 2003 · I have found in Machinery Hndbk where the shear strength for steel is 0.75*Su. I also have peers who swear by the criteria of 0.6*St. My old ...Ultimate Shear Stress - Eng-TipsClarification of bolt strength - Eng-TipsMore results from www.eng-tips.com
  26. [26]
    An investigation into the combined effect of strain rate and ...
    Jul 21, 2025 · The results showed that shear strength of the material decreases linearly with temperature and that corrected strength increases ...
  27. [27]
    The Influence of Strain-Rate History and Temperature on the Shear ...
    The results are discussed in terms of plastic flow mechanisms involved, thermal activation, changes in microstructure, strain ageing and dynamic recovery.<|control11|><|separator|>
  28. [28]
    Shear Yield Stress - an overview | ScienceDirect Topics
    For steels the shear yield stress is usually taken as equal to 0.75×yield stress in tension [4]. A typical choice of material would be a nickel–chromium ...
  29. [29]
    [PDF] Chapter 9. Shear and Diagonal Tension
    Large V (shear force), Large M (bending moment). Formation of flexure cracks precedes formation of shear cracks. Flexure- shear Crack. Flexure-Tension Crack v ...
  30. [30]
    [PDF] Shear Strength of Reinforced Concrete Beams per ACI 318-02
    Stirrups and other shear reinforcement must extend to a distance d from the extreme compression fiber and shall be adequately anchored at both ends according ...
  31. [31]
    Punching Shear Capacity Calculator | SkyCiv Engineering
    The failure "cone" through which punching shear occurs is referred to as the "critical shear perimeter". This perimeter is proportional to the dimensions of the ...
  32. [32]
    [PDF] EN 1992-1-1: Eurocode 2: Design of concrete structures
    6.2.2 Members not requiring design shear reinforcement. 6.2.3 Members requiring design shear reinforcement. 6.2.4 Shear between web and flanges. 6.2.5 Shear at ...
  33. [33]
    [PDF] Strengthening concrete members in shear - Hilti
    Effectively resisting shear necessitates the addition of specific shear reinforcement – known as stirrups, links, or ties – that will activate after the ...
  34. [34]
    Web Stiffeners - A Beginner's Guide to Structural Engineering
    Transverse stiffeners for shear (ie web stiffeners) are used to increase shear capacity in a beam web. This is accomplished by providing out of plane stiffness.
  35. [35]
    SAP2000 Features | STRUCTURAL ANALYSIS AND DESIGN
    From its 3D object-based graphical modeling environment, this intuitive interface allows you to create structural models quickly without long learning curve ...
  36. [36]
  37. [37]
    The Iosipescu shear test method as used for testing polymers and ...
    The Iosipescu shear test, developed initially in the 1960s for use with homogeneous, isotropic structural materials such as metals, has been adapted during ...
  38. [38]
    D3518/D3518M Standard Test Method for In-Plane Shear Response ...
    Nov 28, 2018 · This test method determines the in-plane shear response of polymer matrix composite materials reinforced by high-modulus fibers.
  39. [39]
    [PDF] STRESS-STRAIN CURVES
    Stress-strain curves are an extremely important graphical measure of a material's mechanical properties, and all students of Mechanics of Materials will ...
  40. [40]
    [PDF] Instrumented Charpy Tests at 77 K on 316L Stainless Steel Welded ...
    Besides impact toughness, the availability of instrumented Charpy data has also allowed deriving estimates of dynamic yield strength and shear fracture.