Fact-checked by Grok 2 weeks ago

Split-Hopkinson pressure bar

The Split-Hopkinson pressure bar (SHPB), also known as the Kolsky bar, is an experimental apparatus used to characterize the mechanical behavior of materials under high strain-rate loading conditions, typically in the range of $10^2 to $10^4 s^{-1}. It consists of three main elastic bars—a bar, an incident (input) bar, and a transmitter (output) bar—made from high-strength materials like , with a small specimen sandwiched between the incident and transmitter bars. Upon , the striker bar generates a wave that propagates through the incident bar, deforms the specimen plastically, and transmits partially to the output bar; strain gauges mounted on the bars measure the incident, reflected, and transmitted pulses to derive the specimen's stress- via one-dimensional wave propagation theory. This setup enables precise evaluation of dynamic properties such as yield strength, hardening, and failure modes under rapid deformation simulating impacts or blasts. The technique traces its origins to Bertram Hopkinson's 1914 development of a single pressure method for quantifying transient pressures from high- detonations or impacts, using transfer in a to infer . In the late , as part of post-World War II research on material response to loading, R.M. Davies improved the setup for better wave measurement, while Herbert Kolsky introduced the split- configuration in to isolate short-duration deformation in small specimens, transforming it into a tool for high-rate . Kolsky's innovation addressed limitations of the original by separating input and output signals, allowing accurate strain-rate control and equilibrium validation in the specimen after several wave reflections. Since its establishment, the SHPB has become a for dynamic materials testing across disciplines, applicable to metals, polymers, composites, rocks, , and even biological tissues in , , torsion, or configurations. Key assumptions include uniform one-dimensional wave propagation, negligible dispersion in the bars, and equilibrium in the specimen, though modifications like or miniaturized bars extend its use to ultra-high rates exceeding $10^5 s^{-1} or softer materials. Applications span , automotive , resistance, , and validation of constitutive models for finite element simulations of or high-speed events.

History and Development

Origins and Invention

The Split-Hopkinson pressure bar originated from the pioneering experiments of Bertram Hopkinson, a professor of mechanism and applied mechanics at the University of Cambridge. In 1914, Hopkinson developed a method to quantify transient pressures generated by high-explosive detonations or bullet impacts, addressing the limitations of existing pressure gauges that could not capture short-duration pulses. His setup involved a long elastic steel bar, suspended as a ballistic pendulum, where an impact or explosion at one end generated a compressive stress wave that propagated along the bar. The free end of the bar was attached to a massive anvil, and the momentum transferred upon wave arrival allowed calculation of the incident pressure pulse's magnitude and duration through measurements of the bar's velocity change. These experiments focused on practical applications such as pressure measurement in gun barrels during projectile impacts, providing qualitative insights into wave propagation in solids under high-speed loading. Hopkinson's pressure bar concept proved instrumental in early 20th-century research, particularly for applications during , where it aided in evaluating performance and armor . However, the original design primarily measured pulse characteristics at a single point and did not facilitate direct material property assessment under dynamic conditions. In 1948, R.M. Davies improved Hopkinson's single-bar technique by introducing electrical methods to measure the relation between pressure and time in high-pressure, short-duration experiments, enhancing the accuracy of wave profile recording. In the late 1940s, following , Herbert Kolsky, who had conducted wartime military research at , adapted Hopkinson's single-bar technique to enable quantitative dynamic testing of material stress-strain behavior. Kolsky introduced a "split" configuration by sandwiching a thin specimen between an incident bar and a transmitter bar, with the impact-generated wave passing through the specimen to produce measurable strains in both bars. This modification allowed for the determination of stress, strain, and in the specimen at high loading rates, up to $10^3 s^{-1}, by analyzing the incident, reflected, and transmitted waves using strain gauges or similar detectors. Kolsky detailed this apparatus and its foundational experiments in his 1949 paper, marking the formal invention of the split-Hopkinson pressure bar (SHPB) as a tool for high strain-rate materials characterization. The early SHPB found use in post-war to study the mechanical response of metals, polymers, and explosives under impact conditions relevant to munitions and protective structures. These applications underscored the technique's value in ordnance development, where understanding transient material behavior was critical for improving projectile design and blast resistance.

Evolution and Key Milestones

In the 1950s and 1960s, the Split-Hopkinson pressure bar (SHPB) saw key standardization efforts led by researchers such as E.D.H. and S.C. Hunter, who established theoretical criteria for valid dynamic testing, including conditions for one-dimensional wave propagation and equilibrium in the specimen. Their work emphasized corrections for and effects, enabling more reliable measurements of high strain-rate material behavior. Concurrently, improvements in bar materials, such as high-strength alloy steels (e.g., with yield strengths exceeding 1500 MPa) and aluminum alloys, were adopted to minimize wave dispersion and match impedance with diverse specimens, enhancing the apparatus's versatility for metals and composites. Instrumentation refinements, including refined placements, further supported these advancements by improving signal fidelity during wave propagation. By the 1970s, the integration of acquisition systems marked a pivotal shift, replacing analog oscilloscopes for capturing signals and enabling precise, real-time processing of wave profiles for stress- curve derivation. This transition facilitated higher resolution in data analysis and reduced errors from manual interpretation. In the and , expansions to and torsion configurations broadened the SHPB's applicability, with Harding and colleagues pioneering adaptations for polymeric materials, such as fiber-reinforced composites, through modified end fixtures and control to achieve uniform deformation at strain rates up to $10^3 s^{-1}. These developments, including torsional bars for testing, addressed limitations in uniaxial setups and highlighted rate-dependent viscoelastic responses in polymers. From the 2000s onward, methodological progress included the incorporation of high-speed imaging for full-field strain measurement and visualization of deformation modes, often synchronized with SHPB signals to validate stress equilibrium. Pulse shaping techniques, using copper or rubber discs on the striker bar, became standard to generate ramped incident waves, ensuring constant strain rates and minimizing oscillations for brittle or rate-sensitive materials. In the 2010s, adaptations extended to nanoscale and biological materials, with miniaturized SHPB variants (e.g., micro-bars) for thin films and modified low-impedance setups (using polymeric bars) for soft tissues like hydrogels, as reviewed in studies emphasizing equilibrium challenges in low-strength specimens. Recent milestones as of 2025 include hybrid electromagnetic SHPB systems, which employ LC-circuit-driven strikers for precise, repeatable pulse generation without pneumatic components, improving control over loading rates for rocks and alloys, as demonstrated in versatile configurations.

Theoretical Principles

Wave Propagation in Bars

In slender elastic rods, such as those used in the Split-Hopkinson pressure bar (SHPB), elastic waves can propagate in several modes, including longitudinal (extensional or compressional), torsional (shear), and flexural (bending) waves. Longitudinal waves involve axial compression or extension along the rod's length, torsional waves involve twisting about the axis, and flexural waves involve bending perpendicular to the axis. For long wavelengths relative to the rod's diameter—typically when the wavelength exceeds ten times the diameter—these waves can be approximated using a one-dimensional (1D) theory, neglecting radial and higher-order variations in stress and strain across the cross-section. This 1D approximation is particularly valid for SHPB bars with a slenderness ratio (length-to-diameter) greater than 10, ensuring the pulse duration produces wavelengths much longer than the bar diameter. The speed of longitudinal waves in such rods is derived from the 1D , combining Newton's second law for the motion of rod elements and for small elastic deformations. Consider a small element of the with cross-sectional area A, \Delta z, density \rho, and E. The net force on the element due to stress difference is A E \frac{\partial^2 u}{\partial z^2} \Delta z, where u(z, t) is the axial displacement. By Newton's second law, this equals mass times acceleration: \rho A \Delta z \frac{\partial^2 u}{\partial t^2}. Dividing by A \Delta z yields the \rho \frac{\partial^2 u}{\partial t^2} = E \frac{\partial^2 u}{\partial z^2}. The general is a wave propagating at speed c = \sqrt{\frac{E}{\rho}}, independent of in the 1D limit. For typical SHPB materials like , this yields c \approx 5000 m/s. At interfaces between two rods or a rod and another medium, such as in SHPB setups, incident waves partially reflect and transmit based on acoustic impedance mismatch. The is Z = \rho c, analogous to for wave amplitude. For an incident wave from medium 1 (Z_1) to medium 2 (Z_2), the is R = \frac{Z_2 - Z_1}{Z_2 + Z_1} and the is T = \frac{2 Z_2}{Z_2 + Z_1}, ensuring continuity of stress and at the interface under 1D assumptions. These coefficients determine the partitioning of the incident pulse; for matched impedances (Z_1 = Z_2), R = 0 and T = 1, yielding full transmission with no reflection. In real cylindrical bars, dispersion arises from three-dimensional effects, primarily lateral inertia and Poisson's contraction, as described by the Pochhammer-Chree theory, causing phase velocity to vary with frequency and leading to pulse spreading and oscillatory tails. This geometric dispersion is negligible when the bar slenderness ratio exceeds 10 and the wavelength is much larger than the diameter (e.g., \lambda / d > 10), allowing the 1D approximation to hold for typical SHPB pulse durations of 50–200 μs. For shorter wavelengths approaching the bar diameter, higher modes contribute, distorting the signal and requiring corrections in data analysis.

One-Dimensional Stress Wave Theory

The one-dimensional forms the foundational for analyzing the of materials in the (SHPB), also known as the Kolsky bar, by decomposing the propagating into incident, reflected, and transmitted components. In this setup, a pulse generated by the impact of a travels along the toward the specimen, where it partially reflects back and through the specimen into the . The bars are assumed to remain throughout the test, allowing the and in each bar to be related linearly by : \sigma = E \epsilon, where \sigma is , \epsilon is , and E is the of the . This assumption simplifies the to one-dimensional longitudinal , governed by the derived from and the linear constitutive , with the c = \sqrt{E / \rho} determined by the bar's modulus E and density \rho. The incident \epsilon_i, reflected \epsilon_r, and transmitted \epsilon_t are measured using on the bars, providing the basis for inferring the specimen's dynamic response. At the specimen-bar interfaces, the relies on the condition to ensure that the forces acting on the specimen are balanced, leading to uniform distribution across the specimen under the assumption of one-dimensional wave propagation. The force at the incident bar-specimen interface is F = A_b E (\epsilon_i + \epsilon_r), where A_b is the cross-sectional area of the bar, and the reflected wave contributes with the opposite sign due to the boundary condition at the interface. Similarly, the force at the specimen-transmitter bar interface is F = A_b E \epsilon_t. For , these forces must be equal, yielding the relation \sigma_s = \frac{A_b}{A_s} E (\epsilon_i + \epsilon_r) = \frac{A_b}{A_s} E \epsilon_t, where \sigma_s is the axial in the specimen and A_s is the specimen's cross-sectional area. This assumes that the specimen is thin enough for stress waves to equilibrate rapidly through multiple internal reflections, preventing significant stress gradients. The further assumes uniaxial states in the specimen and bars, neglecting radial and effects at the interfaces. The strain rate in the specimen is derived from the particle velocity discontinuity at the incident bar-specimen , which drives the deformation. The particle velocity associated with the incident wave is v_i = c \epsilon_i, while the reflected wave imparts v_r = -c \epsilon_r; thus, the velocity imparted to the specimen is v_s = c (\epsilon_i - \epsilon_r). For a specimen of L_s, this velocity change corresponds to a uniform \dot{\epsilon}_s = \frac{2 c}{L_s} \epsilon_r, where the factor of 2 accounts for the relative motion between the incident and transmitter bar ends compressing the specimen from both sides (noting the where \epsilon_r is negative for ). This expression assumes uniform deformation throughout the specimen volume, which requires the specimen to be short relative to the of the pulse. The overall in the specimen is then obtained by integrating the over time: \epsilon_s = -\frac{2 c}{L_s} \int_0^t \epsilon_r(\tau) \, d\tau. These relations enable the construction of the specimen's stress-strain curve from the measured wave signals, provided the assumptions hold. The validity of the one-dimensional theory in SHPB testing hinges on several key requirements to ensure the assumptions of uniaxial , uniform deformation, and negligible . The specimen L_s must be sufficiently short such that the round-trip travel time for the wave across the specimen, $2 L_s / c, is much less than the f of the incident , typically L_s < 2 c f / 10 to allow multiple internal reflections for equilibration before the decays. Additionally, between the bars and specimen is crucial; the Z = \rho c A of the bars should be comparable or higher than that of the specimen to promote efficient wave transmission without excessive at the output . The bars must have a large enough relative to their to justify the one-dimensional , avoiding higher-order modes that could cause . These conditions ensure that the measured waves accurately represent the specimen's average response without significant wave or non-uniformity.

Apparatus and Operation

Basic Components

The standard Split-Hopkinson pressure bar (SHPB) apparatus consists of three primary bars: a striker bar, an incident bar, and a transmitter bar. The striker bar, typically launched using a compressed , generates a pulse upon impact with the incident bar. All three bars are commonly constructed from high-strength , such as grade C-350, with a of approximately 200 GPa and a strength exceeding 1 GPa to ensure elastic deformation during testing. These bars usually have a of 12.7 and lengths ranging from 1 to 2 m for the incident and transmitter bars, while the striker bar is shorter, often around 0.3 to 1 m, to control the pulse duration. The test specimen, a thin cylindrical sample with a of 5-8 mm and thickness of 5-10 mm, is positioned between the ends of the incident and transmitter bars to ensure a one-dimensional stress state within the sample. The contact surfaces between the specimen and the bars are lubricated, typically with (MoS₂) or similar agents, to minimize frictional effects and promote uniform deformation. Instrumentation for the SHPB includes semiconductor or foil strain gauges mounted on the incident and transmitter bars, configured in a Wheatstone bridge circuit to measure transient strain signals with high sensitivity. These gauges are positioned 50-100 mm from the specimen interface to capture the incident, reflected, and transmitted waves after dispersion effects are minimized. For tension testing configurations, an optional momentum trap bar may be attached to the end of the incident bar to absorb residual momentum and prevent rebound. Safety features of the apparatus include an enclosed firing chamber for the gas gun to contain potential debris from launches and rigid alignment fixtures, such as precision rails or bearings, to ensure collinear positioning of the bars and prevent misalignment-induced failures. These components collectively facilitate the and of one-dimensional stress waves through the specimen for high strain-rate .

Standard Procedure

The standard procedure for conducting a basic test using the split-Hopkinson pressure bar (SHPB) begins with meticulous preparation of the specimen and apparatus to ensure accurate one-dimensional wave propagation and minimal frictional losses. The specimen is machined to precise dimensions, typically maintaining a length-to-diameter of approximately 1:1 for ductile materials or adjusted for brittle ones (e.g., 0.5:1 to 1:1) to prevent while allowing uniform deformation; the ends must be flat and parallel within tight tolerances, often verified using a micrometer or surface profilometer. Surfaces are then lubricated with a thin layer of or similar to reduce interfacial between the specimen and the incident/transmitter bars. Finally, the bars are aligned collinearly on the mounting frame, with end faces cleaned using wipes and checked for parallelism to avoid misalignment-induced errors in wave transmission. Once prepared, the test is initiated by firing the bar using a gas gun, typically charged with compressed , to accelerate the striker to velocities between 5 and 50 m/s, generating a pulse of 100 to 500 μs duration that propagates through the incident bar to the specimen. The length and gas are adjusted to achieve the desired , with shorter strikers producing higher rates but shorter pulses to prevent wave overlap in the bars. During the test, signals are captured using or strain gauges mounted on the incident and transmitter bars, amplified, and recorded via an or acquisition system at sampling rates of 1 to 10 MHz to resolve the rapid wave transients accurately. The system is synchronized with a velocimeter or similar device to measure the striker's impact independently, ensuring precise of incident wave amplitude and overall timing of the strain traces. Following the test, the specimen is carefully recovered from between the bars for subsequent analysis, such as via scanning electron microscopy to examine failure modes and microstructural changes under high strain rates. The bars are inspected for any surface damage or deformation, and residual gas pressure in the gun is safely released to prepare for the next test.

Testing Configurations

Compression Testing

In compression testing with the split-Hopkinson pressure bar (SHPB), the apparatus is configured to generate compressive stress waves through axial impact of a striker bar onto the incident bar, with the specimen sandwiched between the incident and transmission bars. Setup modifications may include an optional momentum trap attached to the transmission bar to absorb the outgoing wave and prevent unwanted reflections that could distort measurements, particularly for longer-duration tests. Typical strain rates achieved in such tests for metals range from $10^{2} to $10^{4} s^{-1}, enabling characterization of dynamic material behavior under rapid loading conditions. To ensure uniform stress distribution across the short specimen (typically achieving when the difference between specimen ends is less than 10%), techniques are employed by attaching thin discs of soft materials, such as annealed or rubber, to the impact face of the or incident bar. These shapers deform plastically upon impact, ramping up the incident gradually over 50–200 μs, which promotes dynamic and nearly constant rates during deformation. For instance, a 1–2 mm thick disc can extend the sufficiently to allow wave propagation through the specimen without inertial gradients. Common materials tested in compression include ductile metals like aluminum alloys and polymers, where rate sensitivity is evident in the stress-strain response. In aluminum alloys such as 7050-T7451, SHPB tests at 10^3 s^{-1} reveal strengths increasing by 20–50% compared to quasi-static rates, highlighting strain-rate hardening. Polymers like exhibit pronounced rate dependence, with flow stresses rising nonlinearly from 10–100 at 10^3 s^{-1} due to viscoelastic effects, as captured in stress-strain curves showing steeper hardening slopes at higher rates. Validation of test validity relies on checking force equilibrium via the one-dimensional wave relation \epsilon_i + \epsilon_r \approx \epsilon_t, where \epsilon_i, \epsilon_r, and \epsilon_t are the incident, reflected, and transmitted strains, respectively; this ensures the specimen experiences uniform , with deviations indicating non-equilibrium conditions.

Tension Testing

Tension testing in the Split-Hopkinson pressure bar (SHPB), also known as the split-Hopkinson tension bar (SHTB), enables the characterization of material behavior under high strain-rate tensile loading, typically ranging from 10^2 to 10^3 s^{-1}. Unlike testing, which serves as the baseline mode for direct axial loading, tension configurations generate a tensile through wave at the specimen . The setup consists of an incident and a transmitter , both typically equipped with threaded or machined ends to facilitate secure specimen attachment. Specimens are often glued using tabs or clamped via dovetail interfaces to ensure uniform transfer and minimize slippage. A hat-shaped momentum trap is attached to the transmitter to absorb excess wave energy after specimen failure, preventing unwanted bar motion and signal . Pulse generation in the SHTB involves impacting the free end of the incident bar with a launched from a gas gun, producing an initial compressive wave that propagates toward the specimen. Upon reaching the incident bar-specimen interface, where the specimen's is lower than that of the bar, the wave reflects as a tensile , deforming the specimen in while transmitting a portion through to the transmitter bar. This reflected tensile wave drives the high-rate deformation, with durations around 100 µs achievable through design optimizations like configurations. Strain rates up to approximately 1300 s^{-1} have been demonstrated, allowing investigation of dynamic tensile properties in materials such as polymers and metals. For softer materials, aluminum bars (e.g., 7075-T6 , 12.7 mm , 2.4 m length) are preferred due to their lower impedance mismatch, reducing wave and improving signal quality. Key challenges in SHTB testing include preventing specimen under compressive precursors and ensuring that occurs in the gage section rather than at the grips. Buckling is mitigated by using short specimen lengths, typically less than 50 mm, to maintain stability during the initial wave interactions. Grip designs are optimized via finite element analysis to induce in the specimen before yielding in the bar ends, with dovetail geometries reducing stress concentrations and promoting uniform deformation. These solutions enable reliable for stress-strain responses at high rates. An example application is the study of fatigue crack growth in alloy steels, where SHTB setups have been used to measure dynamic and crack initiation under impact loading, revealing rate-dependent increases in toughness values.

Torsion Testing

The torsional variant of the Split-Hopkinson pressure bar, known as the torsional split Hopkinson bar (TSHB), enables the measurement of dynamic properties in materials under high rates by generating and propagating torsional . The setup typically utilizes long bars that are either solid or hollow, with thin-walled tubular configurations preferred to minimize wave dispersion and ensure uniform distribution. The specimen, a short thin-walled (wall thickness to radius ratio t_s / r_s < 0.1), is adhesively bonded or gripped between the incident bar and the transmitter bar. Torsional pulses are initiated through mechanisms such as a rotating flyer releasing prestored momentum or electromagnetic loading devices that rapidly apply without direct . The underlying wave mechanics in the TSHB rely on one-dimensional shear wave propagation, where the shear wave speed is given by c_s = \sqrt{G / \rho}, with G denoting the and \rho the of the bar material. For thin-walled tubular specimens, the applied T relates to the average \tau across the wall by the equation T = 2 \pi r^2 t \tau, where r is the radius of the and t is the wall thickness; this approximation holds under the assumption of uniform shear and negligible radial inertia effects. gauges mounted at 45° to the capture the incident, reflected, and transmitted torsional , allowing derivation of specimen shear stress and via wave impedance matching. TSHB experiments achieve shear strain rates in the range of $10^2 to $10^4 s^{-1}, significantly higher than quasi-static conditions, which facilitates investigation of rate-sensitive behavior. A key advantage is the imposition of loading, free from hydrostatic pressure components that can confound interpretations in axial configurations, thereby providing cleaner data on shear yield strength and without triaxiality influences. Notable applications encompass mapping dynamic yield loci in metals, where TSHB data combined with axial tests delineate the evolution of yield surfaces under high-rate multiaxial loading, as demonstrated in studies on aluminum alloys revealing expanded loci compared to static conditions. Additionally, the technique elucidates shear banding in amorphous materials like bulk metallic glasses, capturing the onset and propagation of localized shear zones at rates exceeding $10^3 s^{-1}, which informs models of instability and fracture in these brittle-yet-ductile systems.

Data Acquisition and Analysis

Strain Measurement

Strain measurement in the Split-Hopkinson pressure bar (SHPB) primarily relies on or strain gauges affixed to the incident and transmitter bars to capture the propagating waves. gauges, with their high gauge factors typically ranging from 100 to 200, provide sensitivities of approximately 100-200 μV/μϵ under standard excitation voltages, enabling detection of low-amplitude transmitted signals in tests involving soft or low-impedance materials. gauges, offering more moderate sensitivities around 2-5 μV/V/μϵ, are widely used for routine high-strain-rate testing due to their robustness and cost-effectiveness. These gauges are mounted in a Wheatstone configuration, often with two active elements positioned diametrically opposite on the bar to double output sensitivity, compensate for axial , and reject common-mode from transverse vibrations. The raw voltage signals from the strain gauges undergo conditioning through dedicated amplifiers featuring bandwidths of at least 100 kHz to faithfully reproduce the rapid transients of stress waves propagating at speeds around 5000 m/s in metallic bars. These amplifiers incorporate low-noise preamplification and optional filters to maintain before . Time-domain traces are then acquired using high-speed digital oscilloscopes with sampling rates exceeding 1 MHz and resolutions of 12 bits or better, ensuring accurate capture of incident, reflected, and transmitted pulses lasting microseconds. For scenarios where contact-based gauges are impractical, alternative non-contact techniques such as laser Doppler vibrometry provide velocity measurements at the bar interfaces, from which strains can be derived via integration, offering immunity to and suitability for remote monitoring. In high-temperature environments exceeding 1000°C, embedded fiber optic sensors like fiber Bragg gratings (FBGs) are integrated into the bars or specimens, leveraging wavelength shifts for strain detection without degradation from thermal exposure. To enhance signal quality, noise reduction strategies include averaging waveforms from multiple identical test shots, which statistically suppresses random electrical and , and applying low-pass filtering with cutoffs around 50-100 kHz to eliminate high-frequency artifacts while preserving the essential wave characteristics. These processed strain signals form the basis for deriving specimen deformation properties.

Stress-Strain Curve Derivation

The derivation of the stress-strain curve in the (SHPB) relies on one-dimensional (1D) wave theory, assuming uniaxial propagation, negligible lateral , and uniform distribution across the specimen after a short . Under these assumptions, the incident, reflected, and transmitted signals—measured via gauges on the input and output bars—are processed to compute the specimen's dynamic response. The at the incident bar-specimen interface is given by v(t) = c (\varepsilon_i(t) - \varepsilon_r(t)), where c = \sqrt{E_b / \rho_b} is the speed in the bars, E_b is the bar , \rho_b is the bar , \varepsilon_i(t) is the incident , and \varepsilon_r(t) is the reflected (typically negative for ). At the specimen-transmitter bar interface, the simplifies to v(t) = c \varepsilon_t(t), with \varepsilon_t(t) the transmitted (positive for ). Equilibrium of forces requires these velocities to be approximately equal during the constant -rate phase, ensuring uniformity in the specimen. The specimen stress \sigma_s(t) is derived from the force balance at the transmitter bar interface, using the one-wave analysis for simplicity: \sigma_s(t) = \frac{A_b E_b}{A_s} \varepsilon_t(t), where A_b is the bar cross-sectional area and A_s is the specimen cross-sectional area. This assumes the transmitted wave carries through the elastic bar, with the specimen deforming plastically or elastically under the applied load. For more accuracy in cases of wave dispersion or impedance mismatch, a three-wave analysis averages the forces: \sigma_s(t) = \frac{A_b E_b}{2 A_s} (\varepsilon_i(t) + \varepsilon_r(t) + \varepsilon_t(t)), but the one-wave form is standard for uniform conditions. The derivation holds for both elastic and plastic deformation regimes, as the stress reflects the transmitted momentum flux under 1D assumptions. The specimen strain rate \dot{\varepsilon}_s(t) follows from the across the specimen length L_s: \dot{\varepsilon}_s(t) = \frac{v_{\text{incident-specimen}}(t) - v_{\text{specimen-transmitter}}(t)}{L_s} \approx -\frac{2 c}{L_s} \varepsilon_r(t), where the approximation uses and the reflected wave dominance during deformation. For , the negative sign of \varepsilon_r(t) yields a positive . The total \varepsilon_s(t), which accumulates including contributions, is obtained by time : \varepsilon_s(t) = \int_0^t \dot{\varepsilon}_s(t') \, dt' = -\frac{2 c}{L_s} \int_0^t \varepsilon_r(t') \, dt'. This accounts for the history-dependent deformation, particularly in regimes where evolves nonlinearly. The resulting \sigma_s(t) versus \varepsilon_s(t) plot yields the dynamic stress- curve at high (typically $10^2 to $10^4 s^{-1}). Raw strain signals are typically processed using numerical tools like , which implement filtering (e.g., low-pass to remove noise), time-shift corrections for wave travel, and the above integrations to generate the curve. Custom scripts or graphical interfaces automate the computation, exporting engineering stress-strain data for analysis. To validate the derivation and apparatus, the high-rate SHPB curves for rate-insensitive materials (e.g., certain metals) are compared to low-rate quasistatic tests, confirming overlap within experimental error and verifying the 1D assumptions.

Limitations and Modifications

Assumptions and Error Sources

The Split-Hopkinson pressure bar (SHPB) technique relies on several core assumptions to derive material properties from stress wave propagation. These include one-dimensional (1D) stress wave propagation within the elastic , which assumes that the is much longer than the to neglect radial effects. Uniform deformation of the specimen is also assumed, implying homogeneous stress and distribution along its length and cross-section during the test. Additionally, frictionless interfaces between the specimen and bars are presumed, ensuring no stresses that could distort axial loading. Adiabatic heating within the specimen is considered negligible, particularly for short-duration tests on metals where is minimal relative to the loading time. Common error sources arise when these assumptions are violated, leading to inaccuracies in measured -strain responses. Inertia effects at the specimen ends can cause non-uniform , resulting in overestimation of stress uniformity. dispersion in the bars, due to higher-order modes, distorts pulse shapes and is typically corrected using Pochhammer-Chree theory to account for frequency-dependent propagation. Misalignment of the bars or specimen, including non-parallel end faces, induces bending modes and uneven distribution, with tight tolerances required to avoid premature . Interfacial further contributes errors by promoting barreling, adding to measured stress through radial constraints. Quantification of these errors is essential for data validity; for instance, non-equilibrium is assessed by the relative difference Δσ/σ, where values exceeding 5% indicate insufficient uniformity and invalidate results from early test phases. Mitigation strategies include for better equilibrium and validation using digital image correlation () to directly measure specimen strains, confirming assumptions like uniformity against bar-derived data. The SHPB is limited for strain rates below 10² s⁻¹, as low loading durations lead to poor signal-to-noise ratios and inability to achieve . It is also unsuitable for very soft materials, such as foams, due to severe impedance mismatch between the specimen and bars, which weakens transmitted signals and amplifies reflection errors.

Advanced Variants

To address the challenges of testing soft materials like foams, gels, and biological tissues, which exhibit low impedance and large deformations, the conventional metallic bars in the Split-Hopkinson Pressure Bar (SHPB) are often replaced with polymeric incident bars to better match the and reduce wave dispersion. This adaptation, known as the Kolsky bar configuration, enables accurate high-strain-rate testing of these materials by minimizing reflections and ensuring one-dimensional . Additionally, viscous pulse shapers—typically made from materials like or — are inserted between the striker and incident bar to tailor the incident , achieving a slower and constant that promotes in the low-strength specimen. These modifications have been particularly valuable in 2010s biomedical applications, such as characterizing the dynamic response of hydrogels and soft tissues for impact injury modeling. High-temperature variants of the SHPB extend testing capabilities to elevated conditions, crucial for alloys in and power generation. Induction heating systems, often using a susceptor coil around the specimen, rapidly heat samples to temperatures exceeding 1000°C while maintaining precise control, with radiation pyrometers providing non-contact temperature monitoring accurate to within 1% up to 1200°C. This setup integrates with the bar assembly via a floating heater to avoid thermal distortion of the bars, allowing dynamic compression tests on materials like under thermal-mechanical loading. Such configurations mitigate heat loss and ensure uniform temperature distribution, enabling the derivation of temperature-dependent curves at strain rates of 10^3 s^{-1}. For combined loading scenarios, triaxial SHPB systems incorporate confining pressure cells to simulate multiaxial stress states, such as those in deep-earth or structural applications. These setups enclose the specimen in a high-pressure chamber (up to 150 MPa) with hydraulic actuators applying radial confinement, while the axial SHPB delivers dynamic compression, revealing enhanced strength and failure modes under superimposed pressures. Miniaturized versions further adapt the apparatus for microscale samples, using bars scaled to millimeters in length and diameter, often integrated with micro-electro-mechanical systems (MEMS) for precise strain measurement and higher strain rates exceeding 10^5 s^{-1}. This allows testing of thin films or microstructures, such as in nanomaterials, by reducing dispersion and enabling sub-millimeter specimen sizes. Post-2020 innovations have focused on enhancing and versatility, with electromagnetic SHPB systems replacing strikers with electromagnetic drivers to generate precise, controllable pulses without physical , achieving rates up to 10^4 s^{-1} with minimal variability across tests. servo-hydraulic SHPB configurations combine electromagnetic or pneumatic loading with servo-controlled hydraulic actuators for intermediate-to-high rates (10^2–10^3 s^{-1}), offering extended pulse durations and for complex loading paths. These advancements have been applied to additive-manufactured materials, such as laser-powder-bed-fusion lattices, where 2023–2025 studies using modified SHPB setups demonstrate anisotropic dynamic strength improvements under , informing design for lightweight structures.

Applications

Material Characterization

The split-Hopkinson pressure bar (SHPB) enables the characterization of rate-dependent mechanical properties of materials under high rates, typically ranging from 10² to 10⁴ s⁻¹, by generating dynamic stress- curves that reveal behaviors not observable in quasi-static tests. Key outputs from SHPB testing include dynamic strength, hardening rates, and failure , which quantify how materials respond to rapid loading. For instance, in magnesium alloys, deformation mechanisms shift from dominant basal slip under quasi-static conditions to profuse twinning at rates around 10³ s⁻¹, leading to increased hardening and altered failure modes. Across material classes, SHPB data highlights distinct rate sensitivities. In metals like steels, rate strengthening occurs due to enhanced dynamics, with strength increasing by approximately 30% at high rates compared to static values, as observed in AISI 1018 under compressive loading (from ~370 MPa to ~480 MPa at ~10³ s⁻¹). Polymers exhibit viscoelastic transitions, where and energy dissipation peak near temperatures, resulting in -rate-dependent and ; for example, polyurethanes show marked stiffening above 10³ s⁻¹ due to molecular chain alignment. In composites, such as carbon fiber-reinforced polymers, SHPB tests identify thresholds under dynamic or , where interlaminar failure initiates at critical rates, influencing overall evolution without full laminate separation. SHPB-derived stress-strain curves serve as essential inputs for calibrating constitutive models that capture , temperature, and strain effects. The Johnson-Cook model, widely used for metals, is fitted using SHPB data to determine strain- hardening parameters (C), as demonstrated in calibrations for high-strength steels where dynamic predictions align closely with experimental curves at 10³ s⁻¹. Similarly, the Zerilli-Armstrong model, which incorporates microstructural influences like , is calibrated via SHPB tests on body-centered cubic metals, improving predictions of softening and in alloys like . Representative examples underscore these capabilities. Aluminum 6061-T6 exhibits a yield strength increase from ~276 quasi-static to ~320 at high strain rates (~3500 ), attributed to strain rate hardening via elevated dislocation densities under rapid deformation. In biological tissues, such as bovine cortical , SHPB compression reveals rate-dependent toughness, with failure strain decreasing by 30-50% at 10³ compared to static rates, informing models for injury thresholds. These characterizations rely on standard SHPB for equilibrium and uniformity validation.

Engineering and Research Uses

In defense applications, the Split-Hopkinson Pressure Bar (SHPB) is employed to simulate ballistic impacts on armor materials, providing high-strain-rate data essential for developing protective equipment and armored vehicles at facilities like the U.S. Engineer Research and Development Center (ERDC). At strain rates up to 20,000 s⁻¹, SHPB tests characterize the dynamic response of metallic armors, ceramics, and composites, generating stress- curves that inform finite element models for . In , SHPB testing evaluates bird-strike effects on composites, such as panels, by measuring dynamic at elevated strain rates to enhance structural against high-velocity impacts. The utilizes SHPB for assessment of energy-absorbing components, testing materials like sheets at rates of 100–1000 s⁻¹ to replicate collision deformations. These experiments validate finite element models by supplying rate-dependent stress- data, improving predictions of energy absorption and structural deformation during frontal crashes exceeding 55 km/h. For instance, modified SHPB setups, including tensile configurations, quantify yield strength variations under , aiding the of alloys and composites for safety. Research frontiers leverage SHPB to investigate explosive welding parameters, where tests at high strain rates determine equations of state coefficients for materials under explosive-driven interfaces, optimizing weld quality and bond strength. In seismic studies, SHPB simulates material responses to blast loading, revealing how rocks exhibit increased peak stress (e.g., up to 246.70 at 14.5 m/s) and fragmentation under dynamic conditions mimicking earthquakes. For additive manufacturing, SHPB assesses defect in 3D-printed lattices, such as auxetic structures from polymeric resins, showing stable at high rates without , which informs defect-tolerant designs. Recent applications include SHPB testing of arc-shaped auxetic structures for enhanced impact (as of 2024). Interdisciplinary applications include biomedical research, where modified SHPB setups with low-impedance bars test soft tissues like porcine at rates up to 4700 s⁻¹ to establish injury thresholds under impact loading. In geomechanics, SHPB evaluates rock behavior under blast loading, demonstrating reduced (e.g., from 25.39 GPa at 25°C to 3.48 GPa at 800°C) and heightened with temperature, critical for underground structure resilience. Overall, SHPB data integrates with simulation tools like for model validation, using raw outputs to replicate high-rate responses in non-metallic materials and refine predictive accuracy for engineering scenarios.

References

  1. [1]
  2. [2]
    Using the split Hopkinson pressure bar to validate material models
    The split Hopkinson pressure bar (SHPB) has been in use for many years to measure the compression response, principally of metals, but also more recently for ...
  3. [3]
    Split Hopkinson Pressure Bar
    May 28, 2024 · The Split Hopkinson Pressure Bar (SHPB), also known as a Kolsky Bar, is a dynamic system designed to test the high strain rate material properties of various ...
  4. [4]
    X. A method of measuring the pressure produced in the detonation ...
    X. A method of measuring the pressure produced in the detonation of high, explosives or by the impact of bullets. Bertram Hopkinson.
  5. [5]
    Highways and Byways in the History of High Rate Mechanical Testing
    Feb 24, 2020 · Hopkinson B (1914) A method of measuring the pressure produced in the detonation of high explosives or by the impact of bullets. Philos ...
  6. [6]
    Review of the development of the split Hopkinson pressure bar
    This paper reviews a range of published literature on the history of various techniques that have been used for the development of a split Hopkinson (Kolsky) ...
  7. [7]
    The dynamic compression testing of solids by the method of the split ...
    Davies and Hunter, 1960. E.D.H. Davies, S.C. Hunter. The dynamic compression testing of solids by the methods of the Split Hopkinson Pressure Bar. (1960). Part ...
  8. [8]
    Split Hopkinson Pressure Bar compression testing of an aluminum ...
    Split Hopkinson Pressure Bar compression testing of an aluminum alloy: Effect of lubricant type ... Davies and S. C. Hunter, J. Mech. Phys. Solids 11 (1963) 155.
  9. [9]
    [PDF] Design of Inserts for Split-Hopkinson Pressure Bar Testing of Low ...
    striker bar, an incident bar and a transmitter bar, all made from 19.05 mm diameter high-strength maraging steel having a nominal yield strength of 2,500 MPa.
  10. [10]
    [PDF] Advancements in the Split Hopkinson Bar Test - VTechWorks
    May 1, 1998 · 1949 Kolsky adds a second pressure bar to Hopkinson's original apparatus, hence the name split Hopkinson bar. Instead of attaching a billet ...
  11. [11]
    John Harding 1983 | PDF | Composite Material - Scribd
    This document summarizes a technique for tensile testing of fiber-reinforced composites at high strain rates, up to 1000 sec-1.Journal Of Materials Science... · A Tensile Testing Technique... · Cfrp ~elostic Specimen' Test...
  12. [12]
    A Review of the Torsional Split Hopkinson Bar - Wiley Online Library
    Oct 10, 2018 · TSHB was first developed by Baker and Yew [19] in 1966. The specimen was designed as a thin-walled tube to maintain the pure shear state.Missing: evolution milestones
  13. [13]
    A split Hopkinson pressure bar for experimental investigation of ...
    Sep 19, 2023 · In order to experimentally investigate rock pulverization, we developed a split Hopkinson pressure bar with compact dimensions and high-speed ...
  14. [14]
    Experimental Methods for Characterizing Dynamic Response of Soft ...
    Jan 27, 2016 · Frew DJ, Forrestal MJ, Chen W (2005) Pulse shaping techniques for testing high-strength steel with a split Hopkinson pressure bar. Exp Mech ...<|separator|>
  15. [15]
    A versatile split Hopkinson pressure bar using electromagnetic loading
    A versatile split Hopkinson pressure bar using electromagnetic loading ... electromagnetic energy conversion technique of LC circuit to generate directly ...
  16. [16]
  17. [17]
    [PDF] AN ANALYSIS OF THE SPLIT HOPKINSON PRESSURE BAR - DTIC
    It is generally conceded that the one-dimensional theory of wave propagation will accurately predict the stresses, strains and particle velocities In long, thin ...
  18. [18]
    [PDF] A Method for Correcting Stress Wave Dispersion in Split-Hopkinson ...
    In conclusion, by presenting the theory behind stress wave propagation in cylindrindical bars, this paper emphasised the importance of dispersion correction for ...
  19. [19]
    TRANSMISSION AND REFLECTION COEFFICIENTS FOR ...
    The complex coefficients of reflection and transmission of elastic waves in rods at the location of a sudden change in cross-section are examined.
  20. [20]
    A Split Hopkinson Pressure Bar Investigation of Impact-Induced ...
    Dec 14, 2023 · A Split Hopkinson Pressure Bar Investigation of Impact-Induced ... All bars are centreless-ground and have nominal diameters of 12.7 mm.
  21. [21]
    Grade 350 Maraging Steel - MakeItFrom.com
    Nov 8, 2019 · Compressive (Crushing) Strength. 750 to 2400 MPa 110 to 350 x 103 psi. Elastic (Young's, Tensile) Modulus. 190 GPa 28 x 106 psi ; Elongation at ...Missing: SHPB | Show results with:SHPB
  22. [22]
    Signal Correction for the Split-Hopkinson Bar Testing of Soft Materials
    Kolsky, H. An Investigation of the Mechanical Properties of Materials at very High Rates of Loading. Proc. Phys. Soc. Sect. B 1949, 62, 676. [Google Scholar] ...
  23. [23]
    [PDF] DESIGN OF A SPLIT HOPKINSON PRESSURE BAR - ABCM
    The apparatus consists of a specimen, the tested material, sandwiched between two elastic bars (see Fig. 1). A right-traveling compressive stress pulse is ...
  24. [24]
    Dynamic Mechanics of Materials Laboratory: Lab Equipment
    The apparatus is comprised of two, 0.875 in diameter 7075-T6 aluminum bars. The incident and transmitter bars are 12 ft.
  25. [25]
    [PDF] Design and Operating Characteristics of a Split Hopkinson Pressure ...
    A split Hopkinson bar apparatus (shown schcmaticall y in Figure 1) has been designed and used to conduct compression tests at strain rates ranizing fror, SO to ...
  26. [26]
    [PDF] A modified split-Hopkinson pressure bar setup enabling stereo ...
    Dec 26, 2022 · The modified setup uses a tube instead of a transmission bar, enabling high-speed cameras and stereo-DIC for non-contact displacement ...
  27. [27]
    [PDF] Instructions for Performing a Split Hopkinson Pressure Bar ...
    After the pressure bars are selected, make sure the end faces of the bars are flat and parallel. 4. Align the pressure bars and striker on the mounting frame.Missing: typical | Show results with:typical
  28. [28]
    Impact loading of additively manufactured metallic stochastic sheet ...
    ... (SHPB) test apparatus. ... To eliminate the impact of air resistance during the test, the target chamber was decompressed using a vacuum pump to a near vacuum.
  29. [29]
    An Investigation of the Mechanical Properties of Materials at very ...
    mechanical behaviour of marerials at very high rates of loading. These are associated with inertia effects in the apparatus and with the recording of transient.
  30. [30]
    [PDF] Pulse Shaping Techniques for Testing Brittle Materials with a Split ...
    Hopkinson pressure bar apparatus is modified by shaping the incident pulse such that the samples are in dynamic stress equilibrium and have nearly constant ...
  31. [31]
    [PDF] Split Hopkinson pressure bar techniques for characterizing soft ...
    Split Hopkinson Pressure Bar (SHPB) is used to characterize dynamic behavior of soft materials, but conventional SHPB is insufficient for accurate stress- ...
  32. [32]
    Split Hopkinson Pressure Bar compression testing of an aluminum ...
    Aug 7, 2025 · The effect of lubricant type on Split Hopkinson Pressure Bar (SHPB) compression testing of an aluminum alloy was investigated.
  33. [33]
    [PDF] Split-Hopkinson Pressure Bar Testing and Constitutive Model ...
    The thesis aims to evaluate the high strain rate response of Aluminum 7050-T7451, IN718 superalloy, and 300M steel, as well as the calibration of constitutive ...
  34. [34]
    [PDF] Using split Hopkinson pressure bars to perform large strain ...
    Dec 25, 2008 · The strain rate sensitivity of polyurea is characterized using a modified split Hopkinson pressure bar. (SHPB) system.
  35. [35]
    Classic Split-Hopkinson Pressure Bar Testing - ASM Digital Library
    This article describes the techniques involved in measuring the high-strain-rate stress-strain response of materials using a split-Hopkinson pressure bar ...<|separator|>
  36. [36]
    [PDF] development of a split hopkinson tension bar for testing stress-strain ...
    Aug 4, 2007 · In the present work, a split Hopkinson tension bar was developed to study high strain rate behavior of polymer composites. The apparatus was ...
  37. [37]
    [PDF] Dynamic Behavior of Ductile Materials - DTIC
    In this study, a tension split Hopkinson bar with a momentum trap is ... special hat-shaped geometry and (b) thin disks deformed in a split Hopkinson ...<|control11|><|separator|>
  38. [38]
    [PDF] Investigations on specimen design and mounting for Split ...
    Hopkinson Bar (SHB) testing methods are often used for this range of strain rates. In these experiments, a specimen is placed between two long elastic rods, the ...
  39. [39]
    Investigation of Dynamic for Alloy Steel Weldments Using the Split ...
    Sep 1, 1986 · The onset of stable crack growth was determined by using a strain gage that was attached ahead of a fatigue crack. The results show that, at −80 ...
  40. [40]
  41. [41]
  42. [42]
    [PDF] A Miniaturized Split Hopkinson Pressure Bar for Very High Strain ...
    This paper describes a miniaturized split Hopkinson pressure bar. (MSHPB) for measuring the stress-strain relationships in materials at strain rates up to ...
  43. [43]
    Review of SHPB Dynamic Load Impact Test Characteristics and ...
    Oct 21, 2023 · The split-Hopkinson pressure bar (SHPB) test technique is based on the Hopkinson pressure bar technique and has undergone a series of ...<|control11|><|separator|>
  44. [44]
    [PDF] Compressive Split Hopkinson Pressure Bar APPARATUS: PART I
    Though Bertram Hop- kinson's observations were qualitative, he was able to correctly establish that the velocity of the impact was the most important cause of ...
  45. [45]
    [PDF] Characterization of a Split Hopkinson Pressure Bar for Measuring ...
    Jul 11, 2024 · During assembly, meticulous attention is required to avoid any deflection and ensure perfect alignment, with supports possessing minimal ...
  46. [46]
    A modified split-Hopkinson pressure bar setup enabling stereo ...
    Furthermore, the method needs a distance from the strain gauge location to the specimen end of the incident bar, which only applies to cylindrical bars and ...
  47. [47]
    Split Hopkinson Pressure Bar | IMPACT | Tampere Universities
    The compression apparatuses built at TAU use high strength 4340 steel (yield strength ~ 1100 MPa), maraging steel (yield strength ~ 1900 MPa), and 7075 aluminum ...<|separator|>
  48. [48]
    Determination of dynamic material properties using laser ...
    Jan 29, 2023 · The split Hopkinson pressure bar (SHPB) is a commonly used technique to measure the stress-strain of materials at high strain rate.
  49. [49]
    Split Hopkinson bar measurement using high-speed full-spectrum ...
    A well-known method for testing material behavior at high strain rates is through a split Hopkinson tensile bar [8,9], which is capable of generating high ...
  50. [50]
    Feasibility of component state awareness of high strain rate events ...
    This paper presents a systematic study that investigates the viability of FBG Sensors under high strain rate loading by initially measuring 1D-strains.
  51. [51]
    [PDF] Influence of filtering size on results in interpretation of Hopkinson ...
    Two methods are used to smooth the signals: low-pass filtering built into the software and arithmetic mean averaging.
  52. [52]
    Split-Hopkinson Pressure Bar Testing
    The Split-Hopkinson Pressure Bar (SHPB) technique complements the Laser Based High Rate Testing by providing a method for deformation at lower rates.
  53. [53]
    Split Hopkinson Pressure Bar Graphical Analysis Tool - MathWorks
    Users can input raw voltages from strain gages and walk through GUI analysis to export stress-strain curves (along with other pertinent data)
  54. [54]
    [PDF] The NIST Kolsky Bar Data Processing System
    A portion of the stress wave is reflected back into the incident bar and a portion travels through the sample and on to the second bar, called the transmitted ...
  55. [55]
  56. [56]
    [PDF] Research progress and discussion on problems of sandy soil SHPB ...
    SHPB data analysis relies on three fundamental assumptions[11]: (1) One-dimensional stress wave propagation. Stress waves undergo dispersion due to lateral ...
  57. [57]
    Influence of strain rate, temperature and adiabatic heating on the ...
    Aug 7, 2025 · At strain rate equal or higher than 0.1 s−1 the thermal softening effect due to adiabatic heating cannot be neglected and a correction procedure ...
  58. [58]
    [PDF] Data-Reduction Uncertainties in Kolsky Bar Experiments on Metals
    Stress wave dispersion, early stress equilibrium, and elastic indentation to the ends of the pressure bars have been accepted as main error sources in stress- ...
  59. [59]
  60. [60]
    An Internal Digital Image Correlation Technique for High-Strain Rate ...
    Oct 23, 2024 · The internal DIC technique has been developed and demonstrated to work well for full-scale dynamic high-strain rate and shock laboratory experiments.Missing: validation mitigation<|separator|>
  61. [61]
    Kolsky Compression Bar Experiments on Soft Materials - SpringerLink
    In this chapter, we present Kolsky-bar techniques modified for soft material characterization in compression. Soft materials typically have low strength, ...
  62. [62]
    Split Hopkinson pressure bar experiments on polymeric foams
    Aug 7, 2025 · Impact responses of extra-soft materials, such as ballistic gelatins and biological tissues, are increasingly in demand. The Kolsky bar is a ...
  63. [63]
    Optimal pulse shapes for SHPB tests on soft materials - Penn State
    Typically, both methods must be employed. Pulse shapes are often tailored to deliver a smooth and sufficiently slow rise to a constant axial strain rate, as ...
  64. [64]
    SHPB experimental method for ultra-soft materials in solution ...
    Since Kolsky [1] proposed the split Hopkinson pressure bar (SHPB) ... soft materials, such as hydrogels and biological tissues [8-12]. However ...
  65. [65]
    Calibration of the Induction Heating System for Dynamic ...
    May 2, 2025 · ... electromagnetic Eddie current excitation. Induction heating was used ... Journal of Dynamic Behavior of Materials 6(3):278–286. https ...
  66. [66]
    [PDF] A SHPB Pulse Shaping Technique for Dynamic Stress-Strain Loop
    Induction heating was required to reach temperatures above 1000°C, which utilized a floating, differential graphite susceptor designed to introduce heat to the ...
  67. [67]
    A technique for measuring the dynamic behavior of materials at high ...
    The most common approach to the measurement of the high-strain-rate response of materials is the compression Kolsky bar or split-Hopkinson pressure bar (briefly ...
  68. [68]
    Developing a microscale split Hopkinson pressure bar
    Mar 10, 2023 · Miniaturization, of both the instrument and samples, is key to achieving higher strain rates. Using a so-called mini-SHPB we have had a ...Missing: miniature MEMS
  69. [69]
    Miniature Beryllium Split-Hopkinson Pressure Bars for Extending the ...
    Oct 28, 2022 · Conventional Split Hopkinson Pressure Bars (SHPB) or “Kolsky” bars are often used for determining the high-rate compressive yield and ...Missing: microscale | Show results with:microscale
  70. [70]
    Electromagnetic Hopkinson bar: A powerful scientific instrument to ...
    Aug 25, 2020 · The split Hopkinson bar (SHB) has been widely used for testing the dynamic mechanical behavior of materials. However, it is hard to involve ...Missing: post- | Show results with:post-
  71. [71]
    [PDF] A modified servo-hydraulic machine for testing at intermediate ... - HAL
    Dec 24, 2010 · In this paper, the hybrid technique of LeBlanc and Lassila [3] is improved by using a wave separation technique. Consequently, the test duration ...
  72. [72]
    High strain rate deformation of additively manufactured materials
    Aug 18, 2025 · In this work, an extensive characterisation and modelling of the high strain rate deformation behaviour of materials fabricated by additive ...
  73. [73]
  74. [74]
    Constitutive relations for slip and twinning in high rate deformations
    Dec 29, 2021 · A strong increase in the strain rate dependence at SHPB rates, compared to conventional tensile or compression test results, had been reported ...
  75. [75]
    (PDF) Deformation Mechanisms and High Strain Rate Properties of ...
    Figure 1. Schematic of the slip and twinning systems in hexagonal close-packed (hcp) Mg. Note that the pyramidal slip is on neither of the two twinning planes.<|separator|>
  76. [76]
    Material characterization at high strain rate by Hopkinson bar tests ...
    Jul 25, 2008 · In the present work, dynamic tests have been performed on AISI 1018 CR steel specimens by means of a split Hopkinson pressure bar (SHPB).
  77. [77]
    High Strain Rate Mechanics of Polymers: A Review
    Jan 25, 2016 · PTFE has been shown to have both strain rate and temperature dependence as expected in polymers. An example of the strain rate dependence is ...
  78. [78]
    Measuring the Effect of Strain Rate on Deformation and Damage in ...
    Feb 15, 2022 · This review focusses on experimental methods for measuring the rate-dependent deformation of fibre composites.Tensile Loading · Compressive Loading · Specimen Geometry And Damage...
  79. [79]
    Determination of Johnson-Cook constitutive model coefficients ...
    In this study, effect of initial gap between specimen and bars was investigated by SHPB test. The process of compressing initial gaps formed by rough places on ...
  80. [80]
    On the performance of modified Zerilli-Armstrong constitutive model ...
    The constitutive data required for calibrating the material constants is ... split Hopkinson pressure bar (SHPB) tests available in literature. The ...
  81. [81]
    Effects of heat treatment and strain rate on the microstructure and ...
    Jan 23, 2015 · As shown in Figure 2, the yield strength σ y of 6061-T6 under quasi-static compression (5 × 10−4 s−1) is 335 MPa. For high strain rates of 2000 ...
  82. [82]
    Constant strain rate compression of bovine cortical bone on the Split ...
    Jan 1, 2015 · Cortical bone is a visco-elastic material which implies that strain rate will affect its response. Although the Split-Hopkinson Pressure Bar ...
  83. [83]
    Split-Hopkinson Pressure Bar & Plate-Impact Testing
    The Split-Hopkinson Pressure Bar at SwRI characterizes dynamic response of materials at high strain rates. Results from these tests are used to generate and ...
  84. [84]
    Numerical Analysis of Bird Impact on a GLARE Plate - Sage Journals
    By using a Split Hopkinson Pressure Bar (SHPB) to perform the material dynamic mechanics experiments at different strain rates, values of five parameters in.
  85. [85]
    [PDF] Material Characterization at High Strain Rates for Optimizing Car ...
    Although split-Hopkinson bar method enables to measure material behavior up to a high strain rate, it is in principle a compres- sion test method. A car body is ...
  86. [86]
    Towards better understanding of explosive welding by combination ...
    May 5, 2019 · It may also be noted, that Split-Hopkinson pressure bar (SHPB) is typically used to determine the coefficient of empirical equations of state.
  87. [87]
    Experimental Study of the Rock Mechanism under Coupled High ...
    Jul 17, 2020 · 1. Development of the Split-Hopkinson Pressure Bar. As for the rock under dynamic loading, sources of dynamic loads could be explosion, impact, ...Missing: biomedical injury
  88. [88]
    [PDF] On the study of the dynamic response in 3D additively manufactured ...
    The Split Hopkinson Pressure Bar (SHPB) was used to test the lattices at a high strain rate by imposing a dynamic load. The diameter of the bars limits the ...Missing: 2023-2025 | Show results with:2023-2025