Fact-checked by Grok 2 weeks ago

Properties of concrete

is a material composed of , , aggregates (fine and coarse), and often admixtures or supplementary cementitious materials, which undergoes a chemical process to form a hard, durable, rock-like suitable for a wide range of structural applications. Its defining properties include high for load-bearing, low tensile strength requiring , good against environmental factors, and versatility in mix design to achieve specific performance characteristics. These attributes make one of the most widely used materials in , from and bridges to dams and . The mechanical properties of concrete are critical for its structural integrity and are primarily governed by the water-to-binder ratio, aggregate quality, curing conditions, and age. Compressive strength, the ability to withstand loads that reduce volume, is the most fundamental property and is typically expressed as the characteristic 28-day strength (f_ck), ranging from 20 for standard mixes to over 60 for high-performance . , which resists pulling forces, is significantly lower at approximately 10% of compressive strength (e.g., 2.9 for a C30/37 mix with 30 compressive strength), making unreinforced brittle under and necessitating reinforcement in most applications. The modulus of elasticity (E_cm), a measure of , is derived from compressive strength and typically varies from 20 to 40 GPa, influencing deformation calculations in design. Physical and durability properties further define concrete's suitability for long-term performance in diverse environments. Density, or unit weight, generally falls between 2200 and 2400 kg/m³, providing substantial mass for stability while allowing for lightweight variants using special aggregates. Durability encompasses resistance to chemical attack, , freeze-thaw cycles, and ingress, enhanced by proper mix design and curing, though recycled aggregates can increase and reduce longevity. Concrete also exhibits time-dependent behaviors such as (gradual deformation under sustained load) and shrinkage (volume reduction leading to potential cracking), which must be accounted for to prevent structural issues. For specialized uses, concrete's properties extend to thermal conductivity (low, aiding ), fire resistance (retaining integrity up to 300°C with ~25% strength loss), and resistance to , shielding, and , all tunable through additives and composition. Workability, the ease of mixing, placing, and compacting, is another key fresh-state property influenced by and admixtures, ensuring constructability without compromising hardened performance. Overall, understanding and optimizing these properties through standards like Eurocode 2 or ACI guidelines enables safe, efficient, and sustainable structures.

Composition and Mix Design

Key Ingredients

Concrete, a , primarily consists of , aggregates, , and admixtures, each contributing uniquely to its structural integrity and performance. acts as the binding agent, while aggregates provide bulk and stability; initiates the chemical reactions necessary for hardening, and admixtures modify specific properties to enhance workability, , or . These ingredients interact during mixing and curing to form a hardened matrix that determines the concrete's mechanical and physical characteristics. Cement, typically , is the core hydraulic binder in , composed mainly of calcium silicates, aluminates, and ferrites produced by heating and clay. When mixed with water, it undergoes —a series of exothermic chemical reactions forming calcium silicate hydrate (C-S-H) and , which bind the mixture into a solid mass. Common types include Type I for general use, Type II for moderate resistance, and Type V for high environments, selected based on exposure conditions to optimize . The process is crucial, as it develops the paste that encapsulates aggregates and imparts strength over time. Aggregates, comprising 60-80% of concrete's volume, serve as the skeletal framework, influencing strength, , and resistance to . They are divided into fine aggregates (e.g., , particles <5 mm) and coarse aggregates (e.g., gravel or crushed stone, particles 5-80 mm), with size distribution optimized for dense packing to minimize voids and enhance load-bearing capacity. The mineralogy of aggregates, such as quartz or limestone, affects durability; reactive types like certain siliceous rocks can cause alkali-silica reactions leading to cracking if not mitigated. Water is essential for initiating cement hydration and achieving workable consistency, but its quantity relative to cement—the water-cement (w/c) ratio—critically governs both fresh and hardened properties. A lower w/c ratio (typically 0.4-0.6) yields higher compressive strength by reducing porosity in the hydrated paste, though it may compromise workability, requiring additional measures for placement. Conversely, higher w/c ratios improve flow and ease of mixing but increase permeability and reduce long-term durability due to excess free water evaporating and leaving voids. Admixtures are auxiliary materials added in small quantities to tailor concrete properties without altering the primary ingredients significantly. Chemical admixtures, such as plasticizers or superplasticizers (e.g., polycarboxylate ethers), enhance workability by dispersing cement particles, allowing lower w/c ratios while maintaining flow for better strength and reduced water demand. Mineral admixtures, often classified as supplementary cementitious materials (SCMs), include fly ash, slag, and silica fume, which exhibit pozzolanic reactivity—combining with calcium hydroxide from cement hydration to form additional C-S-H gel, improving long-term strength and impermeability. Since the 2000s, SCMs like fly ash (a coal combustion byproduct) have seen increased adoption, replacing up to 30-50% of to lower carbon emissions from clinker production and promote sustainability, while also mitigating issues like alkali-silica reaction.

Mix Proportions and Their Impact

Mix proportions in concrete refer to the relative quantities of cementitious materials, water, aggregates, and admixtures used to achieve desired performance characteristics such as strength, durability, and workability. These proportions are determined through systematic mix design processes that balance the need for adequate hydration—where water reacts with cement to form binding compounds—with the goal of minimizing excess water to enhance strength and reduce porosity. The American Concrete Institute (ACI) 211 method provides a standardized procedure for selecting proportions in normal-density, heavyweight, and mass concrete, involving steps like specifying slump for workability, selecting maximum aggregate size, estimating mixing water and air content based on exposure conditions, and calculating cementitious content to meet target compressive strength. This method relies on empirical data tables for water requirements adjusted for aggregate characteristics and environmental factors such as sulfate exposure or freeze-thaw cycles, ensuring the mix suits specific durability demands. The absolute volume method, integral to ACI 211, computes batch quantities by calculating the solid volume of each ingredient—excluding voids—and ensuring their sum equals the total batch volume, typically 1 m³ or 27 ft³; for example, the volume of cement is mass divided by (specific gravity × 1000 kg/m³ for water), with similar calculations for aggregates and water. A key factor in mix design is the water-to-cement (w/c) ratio, which governs compressive strength according to Abrams' law, formulated in 1918: the compressive strength f_c is proportional to \frac{1}{(w/c)^n}, where n is typically 4 to 5 for Portland cement under standard curing conditions, meaning lower w/c ratios yield higher strengths by reducing pores formed during hydration. For normal concrete, typical w/c ratios range from 0.4 to 0.6, selected based on target strength (e.g., 0.5 for 30 MPa concrete) and exposure; ratios below 0.4 often require admixtures to maintain workability. Admixtures, particularly water-reducing agents like superplasticizers, enable reductions in w/c by 10-30% without compromising slump, as they disperse cement particles to improve flow while preserving the dense matrix for enhanced durability. Modern standards have evolved toward performance-based designs, emphasizing verifiable outcomes like long-term durability over prescriptive ratios. Eurocode 2 (EN 1992-1-1:2023), incorporates performance criteria for sustainability and assessment of existing structures, allowing optimized mixes that meet exposure classes (e.g., XA for chemical attack) through factors like minimum cement content and w/c limits tailored to life-cycle performance, as influenced by the fib Model Code 2020's sustainability-driven framework.

Mechanical Properties

Compressive and Tensile Strength

Compressive strength represents the maximum uniaxial compressive stress that concrete can sustain at failure, typically measured on cylindrical specimens loaded along their axis until rupture. For normal-weight concrete used in general construction, this value commonly ranges from 20 to 40 MPa, while high-performance concretes can exceed 100 MPa through optimized mix designs and advanced admixtures. The standard evaluation occurs at 28 days of curing, as this period captures the majority of strength gain under controlled moist conditions, per guidelines. Tensile strength, in contrast, is the maximum stress concrete endures under direct pulling forces before cracking, though direct tension tests are rare due to practical challenges; instead, indirect methods like splitting (cylinder) or bending (beam) tests are employed to estimate it. This strength typically constitutes 8% to 15% of the compressive strength, reflecting concrete's inherent brittleness in tension, where failure occurs abruptly with minimal plastic deformation and often propagates via microcracks. The denoted tensile strength f_t contrasts with compressive notation f_c', underscoring the material's anisotropic behavior under load. High-strength concrete exceeding 60 MPa emerged prominently in the 1980s, enabled by superplasticizers that allowed lower water-cement ratios without compromising workability, surpassing earlier plateaus around 60 MPa from the 1970s. Aggregate interlock contributes to overall strength by providing mechanical shear transfer across crack interfaces in compression, yet the interfacial transition zone (ITZ)—a porous, weaker layer around aggregates—often governs failure by initiating cracks at these boundaries, reducing effective load-bearing capacity. This ITZ vulnerability highlights concrete's composite nature, where matrix-aggregate interactions limit ultimate performance despite high binder strengths.

Elastic Modulus and Deformation

The elastic modulus, often denoted as E or E_c, quantifies the stiffness of concrete and represents its resistance to elastic deformation under applied stress. It is defined through Hooke's law, which relates uniaxial stress \sigma to corresponding strain \epsilon in the linear elastic range as \sigma = E \epsilon. This modulus is a fundamental mechanical property used in structural design to predict short-term, reversible deformations under service loads. Concrete's elastic modulus can be measured via static or dynamic methods, with static tests involving direct loading to determine secant or tangent modulus from stress-strain curves, while dynamic methods use ultrasonic pulse velocity or resonance to compute it from wave propagation. Static moduli are typically lower than dynamic ones due to minor inelastic effects in loading, with differences ranging from 20% to 63% depending on concrete strength and curing conditions; for normal-strength concrete, dynamic values often exceed static by 20-40%. For normal-weight concrete, the elastic modulus generally falls in the range of 20-40 GPa, influenced primarily by compressive strength and mix composition. The elastic modulus correlates strongly with compressive strength f_c', as higher-strength concretes exhibit denser microstructures that enhance stiffness; an empirical relation is E \approx 4700 \sqrt{f_c'}, where E is in MPa and f_c' in MPa. The American Concrete Institute (ACI) 318 standard provides an estimation formula for normal-weight concrete: E_c = 57,000 \sqrt{f_c'}, with E_c in psi and f_c' in psi, applicable for design purposes up to strengths around 70 MPa. Poisson's ratio \nu for concrete, which measures lateral strain relative to axial strain under uniaxial loading, typically ranges from 0.15 to 0.20 for both normal- and lightweight-aggregate concretes in the elastic regime. This low value reflects concrete's quasi-brittle nature and relative incompressibility compared to metals. The linear elastic behavior holds up to approximately 30-50% of the ultimate compressive stress, beyond which nonlinear effects emerge. Aggregate stiffness significantly influences the overall elastic modulus, as aggregates comprise 60-75% of concrete volume and dominate the composite stiffness; stiffer aggregates like crushed stone yield higher E values, while lightweight aggregates (e.g., expanded clay or shale) reduce E by 25-50% compared to normal-weight mixes of equivalent strength due to their lower intrinsic modulus (often 5-15 GPa). Beyond the elastic limit, microcracking initiates at the aggregate-paste interface or within the cement matrix, marking the onset of damage and transition to inelastic response, typically at stress levels above 30-40% of ultimate capacity. This provides context for elastic limits relative to ultimate strengths without delving into failure mechanisms.

Creep Behavior

Creep in concrete is defined as the progressive, time-dependent deformation that occurs under sustained compressive stress after the initial elastic strain. This irrecoverable strain arises primarily from the viscoelastic nature of the hydrated cement paste and is quantified using the creep coefficient φ, which represents the ratio of total strain (elastic plus creep) to elastic strain alone, φ = ε_total / ε_elastic. Typical values of φ range from 1.5 to 3.0 for normal-strength concretes after several years of loading under standard conditions. Basic creep refers to the deformation observed in sealed specimens where no moisture exchange occurs with the environment, while drying creep denotes the additional strain in specimens exposed to drying conditions, often linked to moisture gradients. The underlying mechanisms involve the redistribution of water within the pore structure of the cement paste and the viscoelastic relaxation of the solid skeleton in the hydrated cement matrix. Key influencing factors include the sustained stress level, which remains approximately linear up to about 40% of the compressive strength f_c'; relative humidity, with lower humidity increasing drying creep; and the age at loading, where concrete loaded at younger ages exhibits greater creep due to ongoing hydration. In prestressed concrete design, creep plays a critical role by inducing long-term prestress losses through deformation of the concrete, which relaxes the tension in the tendons and must be accounted for to ensure structural serviceability. Predictive models, such as the updated in the fib Model Code 2010, enable estimation of creep effects by separating basic and drying components and incorporating environmental and material parameters. Refinements in the 2020s, including those in the , enhance these models to address sustainability, such as improved predictions for concretes with recycled aggregates that exhibit higher creep. The total creep compliance function J(t, t_0), which describes strain per unit stress from loading at age t_0 to age t, is expressed as: J(t, t_0) = \frac{1}{E} + \frac{\phi(t, t_0)}{E} where E is the elastic modulus of concrete at the age of loading. A logarithmic approximation for the time evolution of the creep coefficient is: \phi(t, t_0) \approx \phi_\infty \left(1 - e^{-k(t - t_0)}\right) with φ_∞ as the ultimate creep coefficient and k a rate parameter.

Physical Properties

Density and Porosity

The density of concrete, defined as its mass per unit volume (ρ = mass / volume), is a fundamental physical property that influences its structural performance and handling characteristics. Bulk density is typically measured in two primary states: oven-dry, where the sample is dried to remove all moisture, and saturated surface-dry (SSD), where the sample is fully saturated but with no free water on the surface. These measurements account for the material's internal voids and moisture content, with oven-dry bulk density often used as a baseline for mix design calculations. For normal-weight concrete, typical bulk densities range from 2200 to 2500 kg/m³, depending on aggregate type and mix proportions, while lightweight concrete achieves densities of 1400 to 1900 kg/m³ through the use of porous aggregates like expanded clay or shale. The specific gravity of concrete components also contributes to overall density; Portland cement has a specific gravity of approximately 3.15, normal-weight aggregates range from 2.6 to 2.7, and water is 1.0 by definition. These values guide the proportioning of mixes to achieve desired densities, with variations arising from aggregate absorption and void content. Porosity refers to the volume fraction of voids within the concrete matrix, expressed as φ = (V_void / V_total) × 100%, and is classified into total (all voids, including gel pores <10 nm and pores >10 nm) and porosity (larger interconnecting pores that dominate fluid transport potential). Total porosity typically ranges from 5% to 20%, increasing with higher -to-cement (w/c) ratios due to excess forming additional voids during ; for instance, w/c ratios of 0.40 to 0.55 primarily affect porosity without altering intrinsic pores. is measured via saturation methods, such as the sample to achieve full saturation followed by weighing (per ASTM C642), or gas pycnometry, which uses displacement to determine skeletal volume and thus void fraction with high precision for both open and closed pores. Air entrainment intentionally introduces microscopic air voids (typically 4-7% by volume for severe freeze-thaw exposure) to enhance , as specified in durability guides, creating a dispersed that mitigates from formation without significantly altering . In sustainable practices, the use of recycled aggregates introduces density variability, often reducing by 5-10% due to higher inherent from adhered , as highlighted in 2020s guidelines promoting recycled concrete aggregate () in paving mixtures to support principles. This variability necessitates adjusted mix designs to maintain consistent performance under emerging sustainability standards.

Water Absorption and Retention

Water absorption in concrete refers to the process by which water enters the material through its porous structure, primarily via , and can be distinguished between initial surface absorption and total absorption under immersion conditions. Initial surface absorption measures the rate at which is drawn into the outer layer, often assessed over short durations like the first hour, while total absorption involves submerging the specimen for extended periods, such as 24 hours, yielding typical values of 2-6% by mass for normal-weight depending on mix design and curing. The sorptivity test, standardized as ASTM C1585, quantifies this by measuring the cumulative ingress of over time on a sealed specimen with one face exposed, providing a coefficient that indicates durability against moisture-related degradation. The rate follows the empirical relation I = S \sqrt{t}, where I is the per unit area (in mm), S is the sorptivity coefficient (in mm/√s), and t is the elapsed time (in s); this model captures the diffusive rise driven by in interconnected pores. Underlying influences this transport, but focuses on the dynamic uptake capacity. The rise model describes movement as an advancing in pores, governed by the balance of adhesive and cohesive forces, leading to a height h proportional to \sqrt{t} similar to the sorptivity relation, which is critical for predicting ingress in unsaturated . Permeability quantifies the ease of fluid flow through concrete under pressure, differing between intrinsic permeability, which depends solely on the pore geometry and is measured in dry conditions, and apparent (or effective) permeability, which is reduced by partial saturation as water blocks pores. Water retention, influenced by factors like curing regime—such as moist curing that promotes denser hydration products—plays a key role in autogenous healing, where retained moisture enables further cement hydration and calcium carbonate precipitation to seal microcracks up to 0.3 mm wide, enhancing long-term durability. Permeability is modeled by Darcy's law: k = \frac{Q L}{A \Delta P}, where k is the intrinsic permeability (in m²), Q is the flow rate (in m³/s), L is the specimen length (in m), A is the cross-sectional area (in m²), and \Delta P is the pressure difference (in Pa); this applies to steady-state flow in saturated concrete. For low water-to-cement ratios below 0.4 combined with additions (typically 5-10% by cement mass), can be reduced to around $10^{-12} m/s, reflecting a refined structure that minimizes interconnected voids and improves resistance to ingress. Assessment of permeability often uses the rapid penetrability per ASTM C1202, which electrically measures ; the 2019 revision refined specimen preparation and voltage application protocols to better correlate with long-term , with reapproval in 2022 emphasizing for high-performance mixes.

Thermal and Hygroscopic Properties

Expansion, Contraction, and Shrinkage

Concrete undergoes volumetric changes due to variations, primarily manifested as or contraction. The linear coefficient of (α) for typically ranges from 8 to 12 × 10^{-6}/°C, influenced by the type, content, and condition, with an average value around 10 × 10^{-6}/°C. This coefficient quantifies the relative elongation or shortening per unit length per degree change, and the volumetric expansion is approximately the linear value for isotropic materials. The thermal strain (ε_th) is calculated as ε_th = α ΔT, where ΔT is the change, enabling predictions of dimensional in structures exposed to environmental fluctuations. Shrinkage in refers to the volume reduction occurring without external loads, driven by moisture loss or internal reactions, and is categorized into plastic, , and autogenous types. Plastic shrinkage arises from rapid of from the fresh mix surface before setting, leading to surface cracking if not controlled by curing methods. shrinkage results from as moisture migrates from the hardened to the environment, typically developing over months and accounting for the majority of long-term volume loss. Autogenous shrinkage occurs in low water-to-cement ratio mixes due to self-desiccation during , where internal relative humidity drops without external , prominent in high-performance concretes. Chemical reactions can induce delayed expansion, notably the alkali-silica reaction (ASR), where reactive silica in aggregates reacts with alkalis in the cement pore solution to form expansive gels, causing map cracking and serviceability issues over years. Mitigation of ASR expansion has been achieved using lithium-based admixtures, such as , which reduce gel formation by altering the reaction kinetics; their application gained prominence in the 1990s for pavements and structures with reactive aggregates. Total shrinkage strains in generally range from 300 to 800 microstrain (×10^{-6}), varying with mix design, environmental , and member size, with drying shrinkage often contributing 60-80% of the total. Predictive models for shrinkage, such as those in ACI 209, estimate time-dependent strains to inform design; the basic form is ε_sh(t) = ε_sh∞ [t / (a + t)]^b, where ε_sh∞ is the ultimate shrinkage strain, t is age in days, and a and b are constants dependent on and mix proportions (e.g., a ≈ 35 days at 50% relative humidity, b ≈ 0.85). These models integrate factors like ambient relative (reducing shrinkage at higher levels) and specimen size (thicker sections exhibit less shrinkage due to slower drying).

Thermal Conductivity and Diffusivity

Thermal conductivity, denoted as \lambda, quantifies concrete's capacity to transfer heat through conduction under steady-state conditions and is governed by Fourier's law, expressed as \mathbf{q} = -\lambda \nabla T, where \mathbf{q} is the heat flux vector and \nabla T is the temperature gradient. For normal-weight concrete, \lambda typically ranges from 0.8 to 2.0 W/m·K, with values varying based on composition and environmental factors. Aggregate type significantly influences \lambda, as denser aggregates like quartz increase conductivity compared to lighter ones such as expanded clay. Moisture content also plays a key role, with an increase of approximately 70% in \lambda from dry to moist conditions observed as water replaces air in pores, due to water's higher thermal conductivity (approximately 0.6 W/m·K versus air's 0.026 W/m·K). Lightweight concrete, often incorporating porous aggregates, exhibits \lambda values below 1.0 W/m·K, making it suitable for applications in building envelopes. Standard measurements of \lambda follow ISO 8301 for the meter apparatus and ISO 8302 for the guarded method, employing steady-state techniques to ensure accurate determination under controlled conditions. Thermal diffusivity, \alpha = \frac{\lambda}{\rho c_p}, where \rho is density and c_p is specific heat capacity, describes concrete's response to transient heat flow and appears in the heat equation \frac{\partial T}{\partial t} = \alpha \nabla^2 T. For typical normal-weight concrete, \alpha ranges from 0.5 to 1.0 × 10^{-6} m²/s, enabling relatively slow heat propagation that contributes to thermal mass benefits in moderating indoor temperatures. Factors affecting \lambda, such as density and moisture, similarly influence \alpha, with lower values in lightweight variants aiding insulation against rapid temperature fluctuations.

Durability and Degradation

Cracking Mechanisms

Cracking in concrete arises primarily from its low tensile strength relative to , typically 7-12% of the latter, leading to failure under tensile stresses that exceed this capacity. The initiation of cracks often occurs at inherent flaws such as pores or interfaces, where stress concentrations amplify local tensile strains, promoting propagation according to linear fracture mechanics principles. In this framework, the K_I at the crack tip is given by K_I = \sigma \sqrt{\pi a} where \sigma is the applied tensile stress and a is the crack length, with fracture occurring when K_I reaches the material's fracture toughness K_c, typically ranging from 0.5 to 1.5 MPa·m^{1/2} for plain concrete. This low K_c value underscores concrete's brittleness, as cracks propagate rapidly once initiated, often limited by the concrete's tensile strain capacity \varepsilon_{tu} \approx 100-200 microstrain before significant damage. Shrinkage cracking results from volume reduction in the cement paste, either through shrinkage, where moisture loss from capillaries generates tensile stresses, or autogenous shrinkage, caused by self-desiccation in low water-to-binder mixes during early . These mechanisms induce restrained tensile strains, particularly in early-age , leading to random surface cracks if the strain exceeds \varepsilon_{tu}. and shrinkage can contribute to these tensile stresses by amplifying restraint effects in restrained elements. Tension cracking, by contrast, develops under flexural or axial overloads, where bending moments or direct pulls create principal tensile stresses perpendicular to the load direction, forming vertical or inclined cracks along paths. cracking manifests as diagonal fissures under transverse shear forces, initiating at ends or near supports when shear stresses exceed the concrete's shear capacity, often intersecting with tension cracks to form 45-degree patterns. Stress concentrations at microstructural heterogeneities, such as air voids, aggregate-paste interfaces, or initial microcracks, serve as primary initiation sites for all crack types, lowering the effective tensile strength by factors of 2-3 compared to homogeneous materials. propagation is governed by the balance between applied and K_c, with widths controlled to mitigate durability issues; limits of 0.3 mm are commonly specified for moderate exposure to prevent ingress of deleterious agents. Fiber reinforcement, particularly or fibers introduced in 2010s-era standards like those from ACI and fib, enhances control by bridging cracks and distributing stresses, often reducing maximum widths by approximately 50% through increased post-cracking . Environmental factors, notably corrosion of embedded steel reinforcement, induce secondary cracking via expansive rust products that exert internal pressure on the surrounding concrete, accelerating cover splitting and widening existing fissures. This corrosion-driven mechanism is exacerbated by crack-permeable chlorides or carbonation, reducing service life by promoting further ingress and stress buildup at vulnerable sites.

Performance at Elevated Temperatures

Concrete exposed to elevated temperatures, such as during fires, undergoes significant degradation in its mechanical properties, primarily due to thermal-induced physical and chemical changes within its microstructure. The compressive strength typically retains about 70-85% of its ambient value at 400°C (depending on aggregate type), dropping to around 20-30% at 800°C. Similarly, the elastic modulus decreases progressively, often losing 20-30% by 200°C and up to 80% by 600°C, reflecting the softening of the cement matrix and aggregate interfaces. These mechanical losses stem from chemical and physical transformations in the concrete's components. Between 100°C and 200°C, of (C-S-H) occurs, releasing bound water and causing initial increase and strength reduction. At higher temperatures of 600-800°C, decomposition of (CaCO₃) in aggregates releases CO₂, further densifying the structure but exacerbating cracking due to volume changes. Concurrently, rapid heating leads to pore pressure buildup from vaporizing trapped in the low-permeability matrix, which can exceed 5-10 and drive explosive spalling. Explosive spalling is a critical failure mode in dense concretes under fire, where moisture-induced pore pressures cause surface layers to detach violently, potentially reducing cross-sectional area by 20-50% and compromising structural integrity. This phenomenon is moisture-driven and exacerbated in high-strength concretes with low permeability; however, incorporating () fibers at dosages of 0.1-0.2% by volume melts at around 160-170°C, creating micro-channels that relieve pressure and mitigate spalling risk by up to 90%. Design standards like Eurocode 2 (EN 1992-1-2, revised 2023) provide fire performance curves based on temperature-dependent reduction factors for concrete strength and stiffness. The second-generation EN 1992-1-2, published in 2023, includes updates to thermal properties and design methods. Recent advancements in ultra-high-performance concrete (UHPC) demonstrate improved fire resistance, retaining approximately 80% of ambient compressive strength at 400°C due to optimized fiber reinforcement and reduced porosity, outperforming conventional mixes in maintaining load-bearing capacity during moderate heat exposure. Heat penetration rates, influenced by thermal conductivity, further dictate the depth of degradation, with slower diffusion in high-density concretes prolonging internal pressure buildup.

Testing and Evaluation

Mechanical Testing Methods

Mechanical testing methods for involve standardized laboratory and field procedures to evaluate key properties such as , tensile strength, and modulus of elasticity, ensuring consistency and reliability in assessing material performance. These tests are governed by international standards, primarily from the American Society for Testing and Materials (ASTM), which specify specimen preparation, loading conditions, and data interpretation to minimize variability and account for factors like concrete age and curing conditions. Specimen preparation typically includes concrete in molds (cylinders of 150 mm by 300 mm height or 100 mm cubes), curing under controlled and (e.g., 23°C and 95% relative humidity for 28 days), and capping ends to ensure uniform loading surfaces, as improper preparation can introduce errors up to 20% in strength results. Age effects are critical, with strength measurements often conducted at 7, 28, or 90 days, as early-age concrete (under 7 days) exhibits lower stiffness due to ongoing . As of 2025, advancements include a five-minute predictive test for cementitious material performance developed at the University of , enabling faster quality assessment without extended curing periods. The compressive strength test, a fundamental method, uses cylindrical or cubic specimens loaded uniaxially in a compression machine at a controlled rate of 0.15 to 0.35 MPa/s until failure, yielding stress-strain curves that reveal peak load and post-peak behavior. ASTM C39/C39M outlines this procedure for cylindrical specimens, emphasizing alignment to prevent eccentric loading, which could reduce measured strength by 10-15%. Results from these tests provide the basis for quality control in construction, with failure typically occurring via shear planes at 45° to the loading axis. Tensile strength testing addresses concrete's inherent weakness in (about 10% of ), employing indirect methods due to the difficulty of pure axial . The splitting tensile test (ASTM C496/C496M) involves placing a cylindrical specimen horizontally and loading it diametrically at 0.7-1.4 /min, calculating tensile strength from the indirect fracture load using the formula T = \frac{2P}{\pi LD}, where T is tensile strength, P is peak load, L is specimen length, and D is diameter. Direct pull-off tests, used for in-situ assessment, adhere to ASTM C1583 by coring a sample and applying tensile via a pull-off apparatus, measuring strength at the concrete-substrate . Modulus of elasticity, indicating , is determined through static or dynamic methods under ASTM C469/C469M, where static tests involve compressive loading at 0.25 MPa/s to 40% of ultimate strength, deriving the chord modulus from the stress- curve as E = \frac{\Delta \sigma}{\Delta \epsilon} between 0.005% (50 microstrain) and the strain corresponding to 40% of ultimate . Dynamic methods use frequency or ultrasonic techniques for faster evaluation, correlating wave to elastic properties with accuracy within 5-10% of static values. Non-destructive testing (NDT) methods complement destructive tests by allowing in-situ evaluation without specimen damage. Ultrasonic pulse velocity (UPV), per ASTM C597, measures the time for a to travel through the , with v = \frac{L}{t} (where L is path length and t is transit time) indicating density and cracking; values above 4.0 km/s suggest good quality . The rebound hammer test (ASTM C805/C805M) assesses surface by rebounding a spring-loaded , correlating rebound number to via curves, though it is limited to near-surface properties and requires site-specific validation. Recent innovations as of 2025 include thermal imaging systems for rapid strength evaluation of sprayed in tunnels, enhancing on-site and efficiency. Since 2015, correlation () has advanced measurement in mechanical tests, using high-resolution cameras to track surface speckle patterns and compute full-field displacements with sub-millimeter accuracy, reducing reliance on traditional strain gauges and enabling better analysis of heterogeneous failure in . Long-term properties like are evaluated through sustained-load compression tests over months or years.

Thermal and Durability Testing

Thermal testing of evaluates its response to , including , , and fire , to ensure performance in applications ranging from building envelopes to high-temperature environments. These methods quantify heat transfer properties and dimensional stability under controlled thermal loads, aiding in the design of structures that maintain integrity during temperature fluctuations or fire events. Standardized protocols, such as those from ASTM and ISO, provide reproducible results for comparing concrete mixes and predicting in-service behavior. The guarded hot plate method, outlined in ASTM C518, measures steady-state thermal transmission properties of specimens by applying a known across a sample sandwiched between hot and cold plates, with edge guards minimizing lateral heat loss. This technique determines thermal conductivity values typically ranging from 0.5 to 2.0 W/m·K for normal-weight , depending on type and moisture content, and is widely used for assessments in materials. Dilatometry assesses by monitoring linear dimensional changes in prisms or cylinders as increases, using push-rod dilatometers that track with high precision over ranges up to 1000°C. This test, aligned with ASTM E228 principles adapted for cementitious materials, reveals coefficients of around 10 × 10^{-6}/°C for , helping predict stresses from thermal gradients in bridges or pavements. Fire resistance testing exposes concrete elements, such as slabs or columns, to the ISO 834 time-temperature , which simulates a fully developed by ramping furnace temperatures to 945°C at 60 minutes and beyond, evaluating load-bearing capacity, integrity, and insulation for specified durations like 120 minutes. This standard , applied in loaded or unloaded configurations, identifies failure modes including surface cracking and loss of structural function, crucial for certifying tunnel linings or high-rise facades. Durability testing focuses on concrete's resistance to over time, using accelerated simulations to forecast long-term performance under chloride ingress, cyclic freezing, and chemical assaults. These protocols measure parameters like penetration depth and mass loss, informing mix designs for harsh exposures such as structures or sulfate-rich soils. Water retention levels can influence outcomes by altering during these exposures. The Rapid Chloride Permeability Test (RCPT) per ASTM C1202 evaluates chloride ion penetration resistance by applying a 60 V DC potential across a 50 mm thick, 100 mm saturated disk for 6 hours, quantifying charge passed in coulombs to classify permeability from very low (<100) to high (>4000). This electrical conductance method correlates with coefficients, guiding the use of supplementary cementitious materials to enhance zone protection against . Freeze-thaw durability is assessed via ASTM C666 Procedure A or B, subjecting or specimens to 300 cycles of rapid freezing in or at -18°C and thawing at 4-21°C, with durability factored from relative of elasticity or weight loss after cycling. Concretes with adequate (4-7% voids) typically retain over 80% , demonstrating resistance essential for pavements in cold climates. Sulfate attack is determined by ASTM C1012, immersing bars (with aggregate-to-cement ratio of 2.75) in a 5% Na2SO4 solution at 23°C and measuring length changes weekly up to 6 months, where expansions below 0.10% at 6 months indicate high . This test highlights the role of low C3A cements or pozzolans in mitigating ettringite formation and cracking in groundwater-exposed foundations. Accelerated tests, such as those in 12390-12, expose specimens to elevated CO2 concentrations (1-4%) at 65% relative humidity and 20-25°C to simulate decades of atmospheric exposure in weeks, measuring depth via indicator to assess cover depth adequacy for passivation. These methods have gained emphasis in 2020s net-zero standards, promoting low-carbon mixes that balance reduced emissions with enhanced for sustainable . Explosive spalling risk during is evaluated through tests under UNI EN 1363-1, which specifies general requirements for resistance including or standard curves on loaded elements, observing spalling via , depth measurement, and mass loss post-exposure. This protocol, often combined with ISO 834 heating, quantifies pore pressure buildup and moisture migration leading to violent fragmentation, with polypropylene fibers added to mixes reducing incidence by 50-70% in high-strength concretes.

References

  1. [1]
    Properties of Concrete - an overview | ScienceDirect Topics
    Concrete is a versatile and widely used construction material known for its exceptional strength, durability and versatility. It is composed of a mix of cement, ...
  2. [2]
    An Introduction to Concrete and its Strengths and Weaknesses
    Feb 16, 2023 · In this introduction to concrete, we'll discuss the properties of concrete and what makes it such a versatile construction material.An Introduction to Concrete · Components and Properties of...
  3. [3]
    Concrete And Reinforcement Properties For Structural Design ...
    Oct 9, 2024 · The 5 Most Important Concrete Properties · 1. Partial safety factor γ · 2. Compression strength · 3. Tensile strength f · 4. E-modulus E · 5.
  4. [4]
    What is Concrete? Types, Composition & Properties | UltraTech
    Properties of Concrete · 1. Poisson's ratio- · 2. Workability - · 3. Strength - · 4. Durability - · 5. Creep · 6. Shrinkage · 7. Unit weight - · 8. Modular ratio - ...<|separator|>
  5. [5]
    Other important properties of concrete
    The following properties are often important for special purposes: density, fire resistance, radiation shielding, resistance to abrasion, and thermal ...
  6. [6]
    What is Concrete Made Of? Definition & Ingredients
    Concrete is made up of three basic components: water, aggregate (rock, sand, or gravel) and Portland cement.Missing: key | Show results with:key
  7. [7]
    Scientific Principles - Concrete
    Water is the key ingredient, which when mixed with cement, forms a paste that binds the aggregate together. The water causes the hardening of concrete through ...
  8. [8]
    Applications of Cement - American Cement Association
    Cement is the key ingredient in concrete that, when mixed with water, acts as the glue that binds the coarse and fine aggregate together to create concrete.
  9. [9]
    Hydration of Portland Cement
    Portland cement is a hydraulic cement, hence it derives its strength from chemical reactions between the cement and water. The process is known as hydration.
  10. [10]
    Cement & Concrete FAQ - American Cement Association
    Portland cement and portland-limestone cement are the most common types of cement used in concrete. ... Type MH, Moderate Heat of Hydration; Type LH, Low ...
  11. [11]
    Cement hydration - Understanding Cement
    By the process of hydration (reaction with water) Portland cement mixed with sand, gravel and water produces the synthetic rock we call concrete. Concrete ...
  12. [12]
    The Role of Aggregate in Concrete
    Aggregates comprise as much as 60% to 80% of a typical concrete mix, so they must be properly selected to be durable, blended for optimum efficiency, and ...<|separator|>
  13. [13]
    CIVL 1101 - Part 6 - Concrete Agregates - The University of Memphis
    Aggregates generally occupy 70 to 80% of the volume of concrete and therefore have a significant effect on its properties. Strength of concrete and mix design ...
  14. [14]
    CIVL 1101 - Part 13 - Strength - The University of Memphis
    In general, at a constant w/c ratio, higher strength can be obtained by using a leaner mixture. If constant workability is maintained, strength will increase ...
  15. [15]
    Types of Admixtures for Concrete
    Pozzolans are cementitious materials and include natural pozzolans (such as the volcanic ash used in Roman concrete), fly ash and silica fume. They can be used ...
  16. [16]
    Towards Sustainable Concrete: Current Trends and Future ... - MDPI
    Supplementary cementitious materials (SCMs) provide a practical solution for reducing greenhouse gas emissions associated with Portland cement production ...Missing: post- | Show results with:post-
  17. [17]
    End-of-Life Materials Used as Supplementary Cementitious ... - NIH
    Apr 22, 2020 · In recent years, the use of SCMs and/or natural pozzolans has increased in the concrete industry, due to their superior long-term performance.
  18. [18]
    211.1-22: Selecting Proportions for Normal-Density and High ...
    This guide to concrete proportioning provides background infor- mation on, and a procedure for, selecting and adjusting concrete mixture proportions. It applies ...
  19. [19]
    [PDF] EXHIBIT 6 - General Guidelines for Concrete Mix Design - ASCE
    ABSOLUTE VOLUME METHOD​​ In a batch of concrete, the sum of the absolute volumes of cementitious materials, aggregate, fibers, water, solids from admixtures and ...
  20. [20]
    Abrams Law - an overview | ScienceDirect Topics
    The compressive strength is controlled by the ratio between the weight of free water (W) and cement (C) as stated in Abrams' law.
  21. [21]
    Generalization of strength versus water–cementitious ratio ...
    For an average Portland cement concrete cured under normal temperature and moisture, Abrams gave the relationship between compressive strength and water–cement ...
  22. [22]
    Why the Right Water Cement Ratio is Crucial for Your Home?
    For typical concrete mixes, the ratio ranges between 0.40 and 0.60, depending on the required strength.
  23. [23]
    [PDF] Water to Cement Ratios
    A 4000 psi air-entrained exterior concrete should be not be specified with a .40 w/cm. That w/cm ratio will yield 5000 psi concrete; greater than what is ...
  24. [24]
    (PDF) The effects of water-cement ratio and chemical admixtures on ...
    Their addition can allow for the reduction of the water to cement ratio (w/c), without affecting the workability of the mixture.
  25. [25]
    [PDF] Water Reducing Admixtures for Concrete
    WRAs improve concrete durability by lowering the water/cementitious ratio, reducing permeability, increasing air content, and creating a concrete mix with ...
  26. [26]
    [PDF] Contributions of the Future Eurocode 2 for Assessment of Existing ...
    ... Eurocode 2 (FprEN1992-1-1:2022 Design of concrete structures) includes Annex I: Assessment of existing structures, which is informative and covers ...
  27. [27]
    Fib MC2020 - A Sustainability-Driven Performance-Based Approach ...
    The vision for the fib MC2020 is to progress the forward-looking objective of enhancing the sustainability of concrete structures.
  28. [28]
    Performance‐based approach in model code 2020: Working on ...
    Apr 25, 2023 · This paper addresses the principles of the performance-based approach to design and assessment and discusses the manner in which it has been introduced in fib ...Missing: 2020s | Show results with:2020s
  29. [29]
    High-Strength Concrete
    High-strength concrete is typically recognized as concrete with a 28-day cylinder compressive strength greater than 6000 psi or 42 Mpa.
  30. [30]
    [PDF] CIP 35 - Testing Compressive Strength of Concrete - NRMCA
    Concrete compressive strength can vary from 2500 psi (17. MPa) for residential concrete to 4000 psi (28 MPa) and higher in commercial structures. Some ...
  31. [31]
    Standards for 7-day and 28-day strength test results
    According to section 19.2.1.3 of ACI 318-19 the specified compressive strength shall be based on the 28-day test results unless otherwise specified in the ...
  32. [32]
    3. Properties of Concrete - CIVL 1101 - The University of Memphis
    Sep 19, 2022 · Concrete has almost no tensile strength (usually measured to be about 10 to 15% of its compressive strength), and for this reason it is almost ...
  33. [33]
    Bond strength of polymer lightweight aggregate concrete - Uygunoğlu
    Aug 10, 2013 · Concrete is an excellent material when subjected to compressive forces, but the tensile strength of concrete is typically only about 8–15% of ...
  34. [34]
    Ultra-high performance concrete: From fundamental to applications
    Concrete strength reaches its plateau at about 60 MPa in early 1970s believed due to the technological barrier of the existing water reducer.
  35. [35]
    Mesoscale Analysis of the Effect of Interfacial Transition Zone ... - NIH
    Dec 10, 2022 · With the increase in interface transition strength, the compressive strength of the concrete becomes stronger; moreover, the number of cracks ...
  36. [36]
    [PDF] Stress – Strain Relationships
    σ = Eε. This relationship is known as Hooke's Law and was first recorded by Robert Hooke, an. English mathematician, in 1678. Note: Hooke's Law describes only ...
  37. [37]
    High-Performance Concretes, 1989-1994 - FHWA-RD-97-030
    4.2.1 Static and Dynamic Elastic Modulus. The modulus of elasticity is generally related to the compressive strength of concrete. This relationship depends ...
  38. [38]
    Moisture Impact on Static and Dynamic Modulus of Elasticity in ... - NIH
    Jul 27, 2024 · This study aimed to elucidate the moisture impact across two normal-weight structural concretes differing in compressive strength (51.6 and 71.4 MPa).
  39. [39]
    (PDF) Comparison of Static and Dynamic Elastic Moduli in Concrete
    The study finds that static moduli (E_s) are generally lower than dynamic moduli (E_d), with differences reaching 63%, 46%, and 22% for low, medium, and high ...Missing: typical | Show results with:typical
  40. [40]
    Effect of aggregate properties on the strength and stiffness of ...
    Lightweight aggregate properties and water/binder ratio significantly affect concrete strength and elastic modulus. At 18% volume, strength is mainly ...
  41. [41]
    Modulus of Elasticity of Concrete - Civil Engineering
    These formulas are based on the relationship between the modulus of elasticity and concrete compressive strength. ... Ec=4700√f′cMPa(i ...
  42. [42]
    Modulus of Elasticity of Concrete Calculator - Structural Calc
    Apr 15, 2021 · ACI 318-19 specifies two equations that may be used to determine the elastic modulus of concrete. When using Equation 19.2.2.1.a, the density of ...
  43. [43]
    Poisson Ratio - an overview | ScienceDirect Topics
    Generally, the range of Poisson's ratio (μ) is 0.15–0.20 for lightweight and normal-weight aggregate concretes when determined from strain measurements taken ...
  44. [44]
    Nondestructive Investigation of Stress‐Induced Damage in Concrete
    Jun 29, 2010 · At higher stress levels (typically beyond 30% of the strength), microcracks start to develop within the concrete mass. Since microcracking is an ...
  45. [45]
    Mechanical Properties of Lightweight Aggregate Concrete ...
    Sep 13, 2022 · Owing to the low modulus of elasticity (Ec) of the aggregate, Ec of LWAC was 25–50% lower than that of NWC with the same compressive strength.
  46. [46]
    Comprehensive Database for Concrete Creep and Shrinkage
    Jul 1, 2015 · Bažant, Z. P., and Baweja, S., “Creep and Shrinkage Prediction Model for Analysis and Design of Concrete Structures: Model B3,” Adam Neville ...
  47. [47]
    [PDF] CHARACTERIZATION OF CONCRETE CREEP AND SHRINKAGE ...
    The CEB MC 2010 model for predicting creep and shrinkage was introduced in the CEB-FIP. Model Code in 2012. This model is very similar to the CEB MC 90-99 model ...
  48. [48]
    [PDF] Bazant, ZP (2001). “Creep of Concrete.”
    The creep, along with the instantaneous (or initial elastic) strain, is defined as the strain difference between loaded and load-free specimens. Exposed to ...
  49. [49]
    Concrete creep at variable humidity: constitutive law and mechanism
    Aug 7, 2025 · Two main mechanisms have been proposed to explain the origin of the drying creep of concrete. ... Zdenek P Bazant · Spencer T. Wu.
  50. [50]
    Prediction of Long-Term Prestress Losses
    In this paper, the authors present an analytical method to predict the long-term prestress losses in precast, pretensioned or post-tensioned concrete members.
  51. [51]
    [PDF] MODEL CODE 2010 CREEP AND SHRINKAGE ... - UPCommons
    This study extends Model Code 2010 to predict shrinkage and creep of recycled aggregate concrete (RAC), which has higher shrinkage and creep than natural ...
  52. [52]
    Creep of recycled aggregate concrete: Experimental database and ...
    Jan 20, 2019 · When predicting the creep coefficient of RAC using the fib Model Code 2010's creep prediction model, relative to the model's performance on ...
  53. [53]
    [PDF] Linear Logarithmic Model for Concrete Creep - DiVA portal
    Abstract. Control of thermal cracking in young concrete is of great importance to ensure a desired service lifetime and function of a structure.Missing: approximation ≈ kt
  54. [54]
    [PDF] Aggregates for Concrete, Chapter 5 - ResearchGate
    The relative density of an aggregate may be deter- mined on an ovendry basis or a saturated surface-dry (SSD) basis. Both the ovendry and saturated surface-dry ...
  55. [55]
    The Effect of Aggregate Properties on Concrete
    The most common classification of aggregates on the basis of bulk specific gravity is lightweight, normal-weight, and heavyweight aggregates. In normal concrete ...Missing: typical | Show results with:typical
  56. [56]
    Aggregates for Concrete
    4 Absolute volume calculations—To calculate the absolute volume an aggregate occupies in concrete, the mass of aggregate is divided by the absolute density ...
  57. [57]
    [PDF] INFORMATION TO USERS
    Addition of water in the normal range of w/c ratio for concrete (0.40 to. 0.55) affects only the capillary porosity. Since the gel porosity is an intrinsic ...
  58. [58]
    Characterization of Porous Cementitious Materials Using ... - NIH
    Jul 12, 2020 · The most convenient and widely used method is the water absorption test, which measures the difference between the saturated and oven-dried ...
  59. [59]
    A comparative study for determining pore volume of concrete
    Aug 8, 2025 · The study reveals that MIP yields a better correlation with strength and pore volume of concrete, than Helium gas pycnometry and water absorption technique.
  60. [60]
    [PDF] 201.2R-01 Guide to Durable Concrete
    ACI 212.3R lists the air-void characteristics required for durability. ASTM ... Freeze-Thaw Resistance of Aggregates in Air-Entrained. Concrete (Powers ...
  61. [61]
    [PDF] Use of Recycled Concrete Aggregate in Concrete Paving Mixtures
    Sep 1, 2022 · Recycled concrete aggregate (RCA) is made from crushed concrete and can replace virgin aggregate in concrete paving mixtures, often from  ...
  62. [62]
    Assessing of water absorption on concrete composites containing fly ...
    Based on the results from Fig. 7 it could be stated that water absorption by immersion were 5.0 %, 5.3 % and 4.6 % for test concrete specimens FA-00, FA ...
  63. [63]
    C1585 Standard Test Method for Measurement of Rate of ... - ASTM
    Sep 22, 2020 · This test method is used to determine the rate of absorption (sorptivity) of water by hydraulic cement concrete by measuring the increase in the mass of a ...
  64. [64]
    Capillary rise test | Concrete Society
    Jun 3, 2025 · A capillary rise test measures the height or weight of water absorbed by a concrete sample when one surface is in contact with water. Sides are ...
  65. [65]
    Towards the estimation of concrete intrinsic permeability
    Unlike intrinsic permeability, which is influenced by concrete's pore structure only, effective permeability is affected by the free moisture content in ...
  66. [66]
    Autogenous Healing in Cementitious Materials with Superabsorbent ...
    Jan 20, 2020 · A possible solution for this problem is the use of superabsorbent polymers (SAPs). These materials are able to adsorb moisture and to absorb ...
  67. [67]
    An Experimental Assessment of the Water Permeability of Concrete ...
    Dec 10, 2020 · Most direct methods of determining the permeability of porous materials are based on Darcy's law [4]. The permeability measurement of concrete ...
  68. [68]
    Estimation of the permeability of silica fume cement concrete
    Silica fume reduces concrete permeability by creating a more impermeable pore structure, especially beyond 8% replacement, making the microstructure denser.
  69. [69]
  70. [70]
    Thermal Expansion and Contraction
    An average value for the coefficient of thermal expansion of concrete is about 10 millionths per degree Celsius (10x10-6/C), although values ranging from 7 to ...Missing: scholarly | Show results with:scholarly
  71. [71]
    A multiscale model for predicting the coefficient of thermal ...
    Nov 10, 2021 · The coefficient of thermal expansion (CTE) is one of the most important physical parameters of concrete. It is defined as the change in the unit ...
  72. [72]
    Prediction of the Thermal Coefficient of Expansion of Concrete
    Apr 1, 1977 · Abstract: A review of the literature identifies age, moisture content, and proportion of the mix as the principal factors affecting the thermal ...
  73. [73]
    [PDF] TIP 17 - Drying Shrinkage of Concrete - NRMCA
    Plastic, autogenous, and carbonation shrinkage are other forms of volume change that occur in concrete. Plastic shrinkage occurs when the top surface of a slab ...
  74. [74]
  75. [75]
    The Use of Lithium to Prevent Or Mitigate Alkali-Silica Reaction in ...
    One of the most efficient means of controlling ASR in concrete containing reactive aggregates is the appropriate use of supplementary cementing materials (SCM).
  76. [76]
    [PDF] The Use of Lithium to Prevent or Mitigate Alkali-Silica Reaction in ...
    Approximately 10 years later, it was discovered that lithium compounds can be used to control expansion due to ASR. There has recently been increased interest ...
  77. [77]
    [PDF] 209R-92 Prediction of Creep, Shrinkage, and Temperature Effects in ...
    Shrinkage is conveniently expressed as a dimensionless strain (in./in. or m/m) under steady conditions of relative humidity and temperature.<|separator|>
  78. [78]
    Report on Factors Affecting Shrinkage and Creep of Hardened ...
    It is common to describe shrinkage in microstrain or millionths, as the value of strain × 106. Thus, 1000 microstrain is equivalent to 1 × 10–3 mm/mm. Fig.
  79. [79]
    Quantification of factors influencing the thermal conductivity of ...
    In conventional concrete and mortar, the thermal conductivity is known to be affected by the type of pozzolanic admixture [7], the amount of moisture, aggregate ...
  80. [80]
    Factors affecting the thermal properties of concrete and applicability ...
    Thermal conductivity increases with an increase in moisture content and it has been reported that thermal conductivity of concrete in moist state is 70% more ...
  81. [81]
    Thermal Conductivity in Concrete Samples with Natural and ... - NIH
    Feb 8, 2024 · The typical range of thermal conductivity for lightweight concrete is 0.2 to 1.9 W/m°K, whereas for conventional concrete, it is 0.6 to 3.3 ...
  82. [82]
    [PDF] Possibilities of Determination of Thermal Conductivity of Lightweight ...
    Presently, the thermal conductivity is determined most frequently using the stationary plate method in accordance with ISO 8301 and. ISO 8302. Its ...
  83. [83]
    Graphene concrete: Recent advances in production methods ...
    The findings indicate that graphene enhances the thermal properties of hydrated cement, which can potentially reduce early age thermal cracking and improve ...
  84. [84]
    Review on the mechanism and mitigation of cracks in concrete
    Due to the low tensile strength of concrete, residual tensile stress builds in concrete at a young age, which may lead to restrained shrinkage cracking.
  85. [85]
    [PDF] Cracks and Crack Control in Concrete Structures
    In this paper, causes of concrete cracking are dis- cussed, including tensile strength of concrete, temperature, shrinkage and creep effects. Recommended crack.
  86. [86]
    [PDF] Determination of fracture toughness in plain concrete specimens by ...
    Jul 1, 2019 · The average value of the critical stress intensity factor for the C0.7 concrete is 1.05 MPa.m1/2. KEYWORDS. Toughness, R-curve, CMOD, SENB, ...
  87. [87]
    Direct tension test and tensile strain capacity of concrete at early age
    Aug 8, 2025 · The mean fracture strain is ultimately approximately 100 microstrain ... Tensile strain capacity of concrete under various states of stress.
  88. [88]
    Influence of Polypropylene and Steel Fibers on the Performance and ...
    Sep 24, 2021 · Additionally, the use of polypropylene fibers has proven its ability to reduce the sensitivity of concrete to cracking due to shrinkage. Sun ...3. Results · 3.2. Compressive Strength · 3.5. Flexural Strength...
  89. [89]
    Environmental effects on cover cracking due to corrosion
    After corrosion initiation, hydrated rust accumulates around the bar, causing pressure and leading to cover cracking. The ability to predict the time to cover- ...Missing: concrete | Show results with:concrete
  90. [90]
    Corrosion-Induced Cracking of Concrete Elements Exposed to a ...
    May 31, 2013 · 1 Cracking of the concrete cover is a serious consequence of corrosion because it accelerates the overall corrosion process. Aside from being an ...
  91. [91]
  92. [92]
    A comprehensive review on the influence of elevated temperatures ...
    Aug 7, 2025 · A significant strength reduction in HPC above 350 °C to 400 °C, with up to a 75% loss in strength observed at 800 °C [73]. At temperatures ...
  93. [93]
    Effect of Elevated Temperature on Mechanical Properties of ...
    Dec 27, 2016 · Elevated temperatures decrease compressive and tensile strength and elastic modulus of concrete, while increasing ultimate strain in ...
  94. [94]
    Effect of High Temperatures on the Microstructure of Cement Paste
    According to Stepkowska, C-S-H gel dehydrates in the temperature range of 110˚C - 450˚C and transforms to C2S and C3S above 600˚C [23] . According to Heikal, ...
  95. [95]
    [PDF] Carbonate aggregate in concrete exposed to high temperatures
    Samples with carbonate aggregates showed significant mass loss due to the decomposition of CaCO3 at temperatures above 600°C.
  96. [96]
    Pore pressure build-up and explosive spalling in concrete at ...
    May 17, 2021 · This paper provides a state-of-the-art review on pore pressure build-up and explosive spalling in concrete materials under elevated temperature.
  97. [97]
    Spalling Prevention of High Performance Concrete at ... - IntechOpen
    As the most effective methods to reduce the risk of explosive spalling, the addition of polypropylene fibers and the use of a thermal barrier are recommended.
  98. [98]
    Reduction of fire spalling of concrete with small doses of ...
    Jul 8, 2021 · The addition of polypropylene (PP) fibres has been shown to reduce the fire spalling propensity of concrete. When including this type of ...
  99. [99]
    [PDF] EN 1992-1-2 Eurocode 2: Design of concrete structures
    It consists in considering the temperature dependent strength reduction factor kp(θ) referring to an ultimate strength value used for calculation, chosen at ...
  100. [100]
    Mechanical and Microstructural Performance of UHPC with ... - MDPI
    At 400 °C, its post-failure strength retains 85% of its original value, and ... NA incorporation improves the splitting tensile strength of UHPC-RA, particularly ...
  101. [101]
    C518 Standard Test Method for Steady-State Thermal Transmission ...
    Sep 28, 2021 · This test method provides a rapid means of determining the steady-state thermal transmission properties of thermal insulations and other materials with a high ...Missing: guarded | Show results with:guarded
  102. [102]
    Thermal Conductivity Measurements of Concrete Using the Heat ...
    Jan 31, 2013 · Thermal conductivity of concrete is measured using the Heat Flow Meter (HFM) and Guarded Hot Plate (GHP) methods, which are standardized ...
  103. [103]
    E228 Standard Test Method for Linear Thermal Expansion of Solid ...
    Jan 9, 2023 · This test method covers the determination of the linear thermal expansion of rigid solid materials using push-rod dilatometers.
  104. [104]
    Dilatometer (DIL) - NETZSCH Analyzing & Testing
    Building materials such as cement and concrete are analyzed using dilatometry to assess dimensional changes during hydration, drying, and thermal cycling— ...
  105. [105]
    Structural test and analysis of RC slab after fire loading
    The fire-resistance test is the thermal loading of an ISO 834 fire curve on one surface of a specimen for 180 minutes. The shape of furnace and the specimen ...
  106. [106]
    Standard Test Method for Resistance of Concrete to Rapid Freezing ...
    Feb 1, 2024 · 1.1 This test method covers the determination of the resistance of concrete specimens to rapidly repeated cycles of freezing and thawing in the ...
  107. [107]
    [PDF] Understanding AASHTO T277 and ASTM C1202 Rapid Chloride ...
    ... Chloride Permeability of Concrete”. • ASTM C1202, “Standard Test Method for Electrical Indication of Concrete's Ability to Resist Chloride Ion Penetration”.Missing: penetrability 2020
  108. [108]
    Freeze-Thaw Resistance of Concrete With Marginal Air Content
    ASTM C 666, "Standard Test Method for Resistance of Concrete to Freezing and Thawing," is the standard laboratory test method for assessing concrete's ...<|separator|>
  109. [109]
    (PDF) Methods of concrete accelerated carbonation test: a review
    Aug 14, 2024 · This review emphasizes the significance of considering these factors when using accelerated carbonation tests to evaluate the durability of concrete and to ...Missing: 2020s | Show results with:2020s
  110. [110]
    [PDF] The GCCA 2050 Cement and Concrete Industry Roadmap for Net Zero
    Oct 12, 2021 · The sector is committed to producing net zero concrete by 2050 and is committed to acting now. reduction in CO2 emissions of 25% associated ...Missing: 2020s | Show results with:2020s
  111. [111]
    Fire resistance test EN 1363-1 - Hagerbach Test Gallery
    Our test facility is accredited by the norm EN 1363-1 and is ideal for: Fire resistance tests on concrete (spalling behaviour); Fire resistance tests on object ...Missing: UNI explosive
  112. [112]
    [PDF] Explosive Concrete Spalling during Large-Scale Fire Resistance Tests
    Apr 5, 2019 · This paper presents a comprehensive investigation of explosive heat-induced spalling observed during a set of large-scale fire resistance tests ...