Open access
Technical Papers
Jan 28, 2022

Flexural Behavior and Design of Ultrahigh-Performance Concrete Beams

Publication: Journal of Structural Engineering
Volume 148, Issue 4

Abstract

Beams made of ultrahigh-performance concrete (UHPC), a fiber-reinforced concrete with high compressive strength and tensile strain-hardening characteristics, exhibit flexural behaviors that are different than those traditionally associated with steel-reinforced conventional concrete beams. These behaviors necessitate the development of new predictive design tools that reflect the effect of the material-level properties on the flexural behavior. The research presented in this paper assessed the flexural behavior of UHPC beams through the displacement-controlled testing of a prestressed UHPC bridge girder to failure. The girder was 18.90 m (62 ft) long and contained a total of 26 17.8-mm- (0.7-in.)-diameter steel strands and no mild steel reinforcement. The testing focused on capturing the intermediate and final behaviors, including first cracking, yielding of strands, moment capacity at the development of a single dominant crack, and rupture of strands. Building on observations from this study and prior research by the authors and others, a flexural design framework, founded on the concepts of equilibrium and strain compatibility, is proposed for beams made with UHPC and reinforced with conventional steel reinforcing bars, prestressing strands, or both. The proposed framework includes considerations to avoid the local straining and subsequent hinging of UHPC beams and to address the ductility of flexural members. The framework is verified by comparing the experimental results of flexural tests performed by the authors and others to the analytical predictions, which predominantly relied on input material parameters obtained from independent material tests.

Introduction

Ultrahigh-performance concrete (UHPC) is an emerging class of cementitious composites, commonly proportioned using particle packing theory and reinforced with discontinuous steel fibers, that offers superior mechanical and durability characteristics compared to conventional and other fiber-reinforced concretes (De Larrard and Sedran 2002; Graybeal 2006a; Habel et al. 2006; Lepech and Li 2009; Bencardino et al. 2010; Magureanu et al. 2012; Wille and Naaman 2013; El-Helou 2016). A typical UHPC-class material exhibits a compressive strength exceeding 124 MPa (18.0 ksi) and a ductile strain-hardening tensile behavior with a cracking strength above 5.0 MPa (0.73 ksi), sustained to a postcracking tensile strain greater than 0.0025 (Russel and Graybeal 2013; Haber et al. 2018; El-Helou et al. 2022). When used in beams, UHPC can offer substantial tensile resistance, reducing the beam’s cross-sectional dimensions, the weight of the superstructure, and the need for traditional steel reinforcement or prestressing strands as compared to conventional concrete (Graybeal 2006b; El-Helou and Graybeal 2019). These features can lower costs by reducing material volumes, load demands, girder depth, labor and construction time, equipment needs, and shipping and handling needs, along with the potential elimination of girder lines or intermediate piers. However, the use of UHPC in primary structural components in the United States has been generally limited due to its increased initial cost and the lack of formal design guidelines that allow for incorporating UHPC’s tensile ductility in design.
Flexural design of concrete members in the United States is founded on mechanical models and sectional design methods specified in the AASHTO Load and Resistance Factor (LRFD) Bridge Design Specifications, henceforth referred to as AASHTO LRFD BDS (AASHTO 2020), and the American Concrete Institute (ACI) 318-19 Building Code Requirements for Structural Concrete, henceforth referred to as ACI 318-19 (ACI 2019). In these documents, the minimal tensile resistance of concrete is ignored, the discrete longitudinal steel reinforcements must be added and proportioned to carry all tensile stresses, and flexural failure is defined at concrete crushing when the compression strain in the concrete reaches 0.003. Moreover, the guidance for flexural effects in AASHTO LRFD BDS (AASHTO 2020) is limited to concretes having a design compressive strength less than or equal to 103 MPa (15 ksi). More information on the flexural design of concrete members according to AASHTO LRFD BDS (AASHTO 2020) and ACI 318-19 (ACI 2019) can be found in the Appendix.
The mechanical behavior of UHPC is significantly different from conventional concrete, necessitating substantive changes in the common understanding of the flexural behavior. Due to the increased UHPC compressive strength, the sustained UHPC postcracking tensile resistance, and the reduced development length of discrete reinforcements embedded in UHPC, beams made with UHPC and reinforced with steel bars or prestressing strands often fail after the formation of a single localized crack, initiated by the pullout of the crack-bridging fibers and followed by the subsequent rupture of the tensile reinforcement (Graybeal 2008; Meade and Graybeal 2010; Yang et al. 2010; Yoo and Yoon 2015; Stürwald 2017; Yoo et al. 2017; Chen et al. 2018; Hasgul et al. 2018; Shao and Billington 2019). In this mode of failure, the peak flexural moment is attained at the initiation of the localized crack, at which point the strain in the extreme tension fiber is equal to the crack localization strain of UHPC (tensile strain limit) and the compressive strains are well below the crushing strain limit. After localization, the loss of the UHPC fiber-bridging resistance mechanism leads to a decrease in the flexural resistance of the member at the localized section, and in many cases, the load and deformation sustained by the member prior to localization will generate a significant increase in curvature at the localized section, leading to the local straining of the tensile reinforcement and the hinging of the beam around the localized crack. These observations accentuate the need for a novel design and analysis framework that accounts for the enhanced mechanical properties of UHPC and captures the new flexural failure modes expected with this class of materials. Using existing conventional concrete specifications to design UHPC beams, in which the concrete tensile resistance is ignored in beam capacity calculations and the flexural strength is governed by reinforcement yielding and concrete compression failure, can result in formulaic simplifications and assumptions that are not consistent with the behaviors of a UHPC member.
Within the international community, a few recommendation or specification documents for the design of UHPC structural members have been published. Of note are the French, Swiss, and Canadian UHPC documents [NF P18-710 (AFNOR 2016a); SIA 2052 (SIA 2016); CSA S6:19 (CSA 2019)]. In each of these documents, the flexural methodology is based on sectional analysis utilizing the concepts of equilibrium and strain compatibility while using new UHPC mechanical models to determine the stresses on strained cross sections. These new mechanical models for UHPC replace or modify existing models for flexural design of conventional concrete members in their respective jurisdictions and account for the UHPC postcracking tensile resistance afforded by the discontinuous fiber reinforcement.
In the French and Canadian design standards, the flexural moment capacity of a steel-reinforced UHPC beam is calculated from the sectional forces at equilibrium corresponding to a linear strain diagram included in a domain bounded by a compression strain limit, corresponding to UHPC crushing, and a reinforcing steel limit, corresponding to the rupture of reinforcing steel. In contrast, the ultimate flexural moment in the Swiss design recommendation is calculated when the tensile strain in the section reaches the UHPC tensile strain limit εt,lim, taken equal to double the value of the crack localization strain εt,loc, as obtained from a material test (i.e., εt,lim=2εt,loc). Fig. 1 presents a comparison of the cross-sectional stress conditions of UHPC members computed following the recommendations of the French, Swiss, and Canadian documents when the strain at the extreme tension layer is equal to the UHPC tensile strain limit εt,lim and with elastic stresses in compression. Note that εt,lim is equal to εt,loc in the French and Canadian models and 2εt,loc in the Swiss model. In Fig. 1, ftcr,F and ftd,F are the design cracking and tensile strength as specified by the French document (Fig. 3.206 of NF P18-710). Similarly, ftd,S and ftd,C are the design tensile stresses computed according to the Swiss (Section 2.4.2.3 of SIA 2052) and Canadian (Fig. CA8.1.8a of CSA S6:19) documents, respectively. The French and Canadian standards allow for tensile strains beyond the UHPC tensile strain limit εt,lim, shown in Fig. 1, in which case the bending moment can be computed when the extreme compression strain is equal to the UHPC crushing limit. In this instance, the contribution of UHPC beyond the tensile strain limit is neglected and the compression stress-strain model is represented by an elastic-perfectly plastic model according to the French standard or a rectangular stress block according to the Canadian recommendation. More information on the UHPC mechanical models and flexural design methodologies in each of the French, Swiss, and Canadian documents can be found in the Appendix. Although these models can predict aspects of the flexural behavior of a UHPC beam, they do not highlight the differences in the behavior before and after crack localization. Because these models allow postlocalization tensile straining of the UHPC, excessive local straining of the tensile reinforcement and hinging of the beam at the localized crack can occur in some instances; these behaviors are not recognized in the design methodologies.
Fig. 1. Cross-sectional stress conditions for UHPC members in flexure computed according to the (a) French; (b) Swiss; and (c) Canadian design documents when the strain at extreme tension layer is equal to the UHPC tensile strain limit and with elastic stresses in compression.
A few additional flexural design models for UHPC members can be found in the literature. A number of these models focus on either developing phenomenological constitutive models utilizing smeared crack finite-element model (FEM) approaches (e.g., Peng and Meyer 2000; Radtke et al. 2010; Cunha et al. 2011; Pros et al. 2012; Li et al. 2019; El-Helou et al. 2020), applying existing damage plasticity models within commercially available FEM software (e.g., Chen and Graybeal 2012; Yin et al. 2019), or formulating lattice or lattice discrete particle model approaches (Bolander et al. 2008; Schauffert and Cusatis 2012; Smith et al. 2014; El-Helou 2016). Although some of these models were validated to capture the behavior of UHPC flexural components accounting for the effect of fiber orientation and tensile ductility, they are not suitable for direct implementation into design guidelines or common design calculation processes.
Realizing the need for simplified analytical procedures, some researchers focused on sectional analysis approaches, applying the concepts of strain compatibility and force equilibrium and proposing new uniaxial stress-strain relationships for UHPC. For instance, authors have suggested that the compression stress-strain curve might be idealized with a linear relationship (e.g., Graybeal 2008; Aaleti et al. 2013), elastic-perfectly plastic relationship (e.g., Soranakom and Mobasher 2009; Yao et al. 2017), or trilinear relationship including an elastic-perfectly plastic and descending branch to failure (e.g., Gowripalan and Gilbert 2000). In tension, the stress-strain curve has been idealized with an elastic-perfectly plastic relationship (e.g., Graybeal 2008; Aaleti et al. 2013), a trilinear relationship with an elastic- perfectly plastic and descending branch to failure (e.g., Gowripalan and Gilbert 2000), or a trilinear relationship with biaxial and plastic branches (e.g., Soranakom and Mobasher 2008). In many of the existing studies, the uniaxial models, particularly the tension models, were obtained from inverse analysis of material testing of small-scale beams subjected to flexure, and the recommendations were limited to the tested specimens and the behaviors of the materials investigated.
The presented considerations indicate the need for unified mechanical models that are inferred from independent uniaxial material testing methods and utilized in a flexural design methodology that is validated with large-scale experimental tests. Such a process is essential to determine whether material-level performance metrics, which are typically available during the design phase, can be relied upon to accurately determine the flexural capacity and failure mechanisms of steel-reinforced beams made with UHPC-class materials. This manuscript addresses this need through the displacement-controlled testing of a full-scale prestressed UHPC bridge girder to failure and the validation of a mechanics-based flexural design methodology that can capture the failure modes of steel-reinforced UHPC beams while relying on material parameters that are available during the design phase. The experimental work focused on capturing specific flexural behaviors at first cracking, yielding of prestressing strands, moment capacity at the development of a localized crack, and rupture of the strands. The flexural design framework is founded on the concepts of equilibrium and strain compatibility and is proposed for beams made with UHPC and reinforced with conventional steel reinforcing bars and/or prestressing strands. The method utilizes existing UHPC mechanical behavior models obtained from uniaxial material tests, accounts for the flexural behaviors and expected failure modes associated with UHPC members, and limits the tensile strain in beam sections to the UHPC strain at crack localization to avoid hinging of the beam around the localized crack. The ductility of UHPC beams is also addressed through a sectional-curvature approach that derives a resistance factor based on a threshold curvature ductility ratio. The flexural design methodology is tailored to generally align with existing bridge design specifications in the United States (AASHTO 2020) in support of ongoing efforts to develop UHPC structural design guidance. The capability of the proposed method to predict the flexural behavior is verified by comparing the experimental data, obtained from the flexural test described in this paper and other tests from the literature, to the analytical predictions, relying on input material parameters obtained from independent material tests whenever available.

Experimental Program

Test Specimen

To conduct the experimental program described in this paper, one pretensioned bulb-tee bridge girder made with UHPC was tested under displacement-controlled loading in flexure to failure. The girder’s cross-sectional shape was based on the Precast Concrete Economical Fabrication (PCEF) Committee recommended sections, with geometrical modifications to reduce the thickness of the bottom flange and the width of the top flange. The standard drawings for the original PCEF sections can be found in the New Jersey DOT Design Manual for Bridges and Structures (NJDOT 2016). The cross-sectional dimensions and reinforcement details of the tested girder are shown in Fig. 2. The girder had a top and bottom flange width of 813 mm (32 in.), a web width of 178 mm (7 in.), and a height of 889 mm (35 in.); it was 18.90 m (62 ft) long and contained a total of 26 17.8-mm- (0.7-in.)-diameter straight steel strands, with 24 strands located in the bottom bulb and two strands in the top bulb, as indicated in Fig. 2. No mild steel reinforcement bars in the longitudinal or transverse directions were included in the girder. The 17.8-mm- (0.7-in.)-diameter strands were Grade 1860-MPa (270-ksi) low-relaxation, seven-wire strands [ASTM A416/A416M (ASTM 2018a)] with a cross-sectional area of 189.7  mm2 (0.294  in.2) and a minimum ultimate strength of 1,860 MPa (270 ksi); they were pretensioned to 75% of their ultimate strength and placed at a minimum spacing of 50.8 mm (2 in.) center-to-center. The first, second, and third layers of strands were located at 50.8 mm (2 in.), 101.6 mm (4 in.), and 838 mm (33 in.) from the bottom of the section, respectively. The gross cross-sectional properties of the girder are an area of 3.72×105  mm2 (577  in.2), a moment of inertia about the centroid of 3.92×1010  mm4 (9.42×104in.4), and a centroid located at 432 mm (17 in.) from the bottom of the section.
Fig. 2. Cross-sectional dimensions and reinforcement details of the tested girder.

Mix Design and Girder Fabrication

The UHPC product used in the girder was proprietary and was supplied in three primary constituents, namely, the preblended powder containing the granular constituents (e.g., cement, silica fume, ground quartz, and fine sand), liquid admixtures, and fibers. The mix design consisted of 2,182  kg/m3 (3,678  lb/yd3) of preblended powder, 166  kg/m3 (280  lb/yd3) of water, 73.4  kg/m3 (124  lb/yd3) of liquid admixtures, and 157  kg/m3 (265  lb/yd3) of brass-coated straight steel fibers (2% dosage by volume). The fibers had a length of 13 mm (0.51 in.), a diameter of 0.2 mm (0.0079 in.), and a supplier-reported minimum tensile strength of 2,600 MPa (377 ksi).
The specimen was fabricated at a precast plant in Florida using one of the plant’s central mixers. In the mixing process, all granular constituents were blended first, then the water and superplasticizers were added gradually. After the mixture turned from a powdered state into a viscous fluid, the fibers were dispersed into the mix. Before casting of each specimen, a flow table test was performed per ASTM C1856 (ASTM 2017) to ensure a flowable consistency corresponding to a spread equal to or greater than 203.2 mm (8 in.).
Two batches of UHPC of approximately 3.44  m3 (4.50  yd3) were mixed to obtain the total volume of fresh UHPC needed to make the girder. The fresh UHPC of the first batch, Batch A, filled the bottom half of the girder’s rigid steel form, while the fresh UHPC of the second batch, Batch B, filled the top half of the form. There was no delay in placement either during or between batches. The placement of fresh UHPC was performed from one stationary discharge point located in the middle of the girder form. The material flowed to each end of the form as it filled the bottom bulb; it then rose upward through the web height before starting to flow again laterally as it filled the top bulb. The discharge point was then moved sequentially along the girder length to top off the filling of the form. Minor external vibration was applied to the sides of the form to facilitate the release of entrapped air. The girder was then covered with a plastic sheet and cured at outdoor ambient temperatures. The placement occurred during the month of June 2019 in which the average daily temperature ranged between 22°C (72°F) and 32°C (89°F). Five days after casting, the forms were removed and the strands were detensioned. The compressive strength required for detensioning was 96.5 MPa (14 ksi); the fabricator elected to wait until the average compressive strength was 130.4 MPa (18.9 ksi). The compressive strength was determined by performing compression tests on two cylindrical specimens (one specimen per batch) having a diameter of 76.2 mm (3 in.) and height of 152.4 mm (6 in.). The cylinders were tested by the precast plant quality control crew according to ASTM C39 (ASTM 2018b) with modifications listed in ASTM C1856 (ASTM 2017).

Test Setup and Instrumentation

The 18.90-m- (62-ft)-long girder was tested in flexure with a span of 18.29 m (60 ft) at the Federal Highway Administration (FHWA) Turner-Fairbank Highway Research Center (TFHRC) structural testing facility. The age of the girder at the time of the test was 155 days after casting. An overview of the test setup is shown in Fig. 3. The girder was supported by a roller support at one end (west side) and by a partial pin support assembly bearing on a load cell and a hydraulic jack at the other end (east side); it was loaded in a four-point bending configuration with the middle reaction points each located at 0.46 m (1.5 ft) from the midspan, as shown in Fig. 4. The roller and partial pin supports each consisted of a 152-mm-(6-in.)-diameter steel cylinder inserted between two steel bearing plates having a thickness of 38.1 mm (1.5 in.), a width of 0.30 m (1 ft), and a length equal to the width of the bottom flange. The bearing plates of the pin support were grooved to prevent rolling, while the plates at the roller support were flat. To relieve frictional forces on the pin support generated by the loading jack line of action relative to the inclination of the bearing location, an additional steel bearing plate resting on a polytetrafluoroethylene (PTFE) sheet was included within the partial pin support assembly, as shown in Fig. 4. When the hydraulic jack applied load and deformation onto the east end of the girder, two reaction points near the midspan resisted the loading. These reaction points consisted of a pair of steel transfer plates reacting against the reaction frame through a set of three load cells with spherical bearing plates (two on the west side and one on the east side of the midspan), a spreader beam, and a greased spherical bearing, as shown in Fig. 3(b). Each of the middle transfer plates had a thickness of 102 mm (4 in.) and a width of 0.30 m (1 ft). All transfer and bearing plates were grouted to the girder’s top or bottom flanges and were long enough to fully support the width of the girder.
Fig. 3. Overview of the experimental test setup showing (a) bridge girder and reaction frame; and (b) instrumentation at midspan.
Fig. 4. Test setup and instrumentation plan.
In addition to the four load cells, the girder was instrumented with six wire potentiometers (WPs) and 13 linear variable displacement transducers (LVDTs). The WPs (WP1–WP6) measured the vertical deflection at selected locations along the span, as shown in Fig. 4. The strain profile throughout the test was captured by mounting a pair of LVDTs, one at the top and one at the bottom of the girder, in five regions (Regions 1–5) covering the middle 2.44 m (8 ft) of the span where monitoring of flexural cracking was desired. Figs. 3(b) and 4 show the top (T1–T5) and bottom (B1–B5) LVDTs mounted on the girder in each of the five regions, with Region 3 being the constant moment region. These LVDTs measured the average longitudinal displacement that can be transformed into average strain by dividing the measured data by the LVDT gauge length. The top LVDTs had a gauge length of 610 mm (24 in.) and were mounted on one side of the top flange at a vertical distance below the top extreme fiber of 25.4 mm (1 in.) for LVDTs T1, T3, and T5, and of 76.2 mm (3 in.) for LVDTs T2 and T4. The bottom LVDTs (B1–B5) had a gauge length of 584 mm (23 in.) and were mounted on the bottom surface of the girder at a vertical distance of 50.8 mm (2 in.) below the girder. Finally, the slippage of the strands during the test was monitored through a set of three LVDTs mounted on three bottom-layer strands on the west end of the girder.
The load was applied by a hydraulic jack in 44.5-kN (10-kip) increments until the girder began to sustain inelastic damage and exhibit a reduced flexural stiffness. The loading was then switched to displacement control at a jack displacement rate of between 1.27  mm/min (0.05  in./min) and 7.62  mm/min (0.3  in./min). Periodically throughout the test, unloading/reloading cycles of 89 kN (20 kip) were performed to measure the residual stiffness of the girder. The loading continued until girder failure, which was defined as the formation of a dominant crack triggered by the pullout of the fiber reinforcement and the subsequent rupturing of the prestressing strands. Given that the girder deformation exceeded the stroke of the loading jack, the test was conducted by loading the girder in three separate test stages. At the end of each stage, the girder was completely unloaded to allow for the necessary preparations to reset the loading apparatus. The first test stage was halted when the load at the jack was 547 kN (123 kips) with a girder vertical displacement (WP6) of 216 mm (8.5 in.). The second stage was halted when the jack load was 812 kN (182.1 kips) with a girder vertical displacement (WP6) of 461 mm (18.1 in.). The third and last stage subjected the girder to a maximum jack load of 854 kN (192 kips), which occurred at a vertical displacement (WP6) of 517 mm (20.4 in.), on its way to complete failure (strand rupture) at a jack load of 761 kN (171 kips) and girder vertical displacement (WP6) of 610 mm (24.0 in.). Note that the stroke of the loading jack was reset once during the second test stage and twice during the third test stage. The reset was performed using additional hydraulic jacks to hold the applied load while the stroke of the loading jack was being retracted and spacer plates inserted between the jack and the girder.

Material Characterization

A key objective of this study was to evaluate the predicted flexural response of the tested girder utilizing a proposed design methodology in which the mechanical properties had been obtained from independent material tests. For this reason, companion specimens were cast from each of the two batches used to make the girder to assess the UHPC’s density, compression, and tension properties at the time of the test. The material specimens were demolded at the time of detensioning of strands and stored in an ambient environment, generally alongside the girder, until the time of the test.
The average results of the material property tests are summarized in Table 1. The average density ω¯c, modulus of elasticity E¯c, compressive strength f¯c, and strain at compressive strength ε¯cu of each UHPC batch were obtained from three cylindrical specimens tested according to ASTM C1856 (ASTM 2017). The cylindrical specimens had a diameter of 76.2 mm (3 in.) and a height of 152.4 mm (6 in.). The average tensile parameters, namely, the effective cracking stress f¯t,cr, the localization stress f¯t,loc, and the localization strain ε¯t,loc, were obtained from prismatic specimens tested in uniaxial tension according to AASHTO T 397 (AASHTO 2022), which is based on the test method by Graybeal and Baby (2013). The tensile test captures the tensile load and associated strain over a 101.6-mm (4-in.) gauge length during a fixed-end, uniaxial displacement-controlled test. The prismatic specimens had a length of 432 mm (17 in.) and a square 50.8  mm×50.8  mm (2  in.×2  in.) cross section. The tensile parameters reported in Table 1 were obtained from the individual stress-strain curves of three and four prisms cast from Batches A and B, respectively. In determining the individual results, the effective cracking stress ft,cr was taken as the stress at the intercept of a line with a slope equal to the elastic modulus and a strain offset of 0.02% (Haber et al. 2018; El-Helou et al. 2022). The localization stress ft,loc and strain, εt,loc, were visually determined as the first point in the stress-strain plot where the stress decreases continuously with increasing strain. This point indicates the onset of fiber pullout and the accumulation of the deformation into a single dominant crack. The average uniaxial stress-strain curves for each batch are shown in Fig. 5(a).
Table 1. Average density and uniaxial compressive and tensile parameters for each UHPC batch at the time of the flexural test
Batch IDDensityCompression parametersTension parameters
ω¯c [kg/m3 (lb/ft3)]E¯c [GPa (ksi)]f¯c [MPa (ksi)]ε¯cuf¯t,cr [MPa (ksi)]f¯t,loc [MPa (ksi)]ε¯t,loc
A2,336 [0.4%] (145.8)43.4 [0.9%] (6,294)173 [2.6%] (25.1)0.00435 [5.0%]9.3 [8.2%] (1.35)10.4 [5.9%] (1.51)0.00497 [13.1%]
B2,323 [0.8%] (145.0)43.7 [2.3%] (6,339)161 [3.1%] (23.3)0.00401 [2.2%]10.2 [3.3%] (1.48)11.6 [7.8%] (1.68)0.00483 [10.0%]

Note: Values in brackets represent the COV, taken as the ratio of the standard deviation to the mean value.

Fig. 5. (a) Individual and average uniaxial stress-strain curves obtained from UHPC tensile tests along with the UHPC tensile response model; and (b) individual uniaxial stress-strain curves obtained from strand tension testing and strands uniaxial response model.
The stress-strain relationships for the prestressing strands were experimentally determined from six strand pullout tests performed according to ASTM A370 (ASTM 2020) and are plotted in Fig. 5(b). The strain data were continuously collected until a uniaxial strain of at least 0.03. The yield strength was measured at 1.0% extension under load and the total elongation value was determined by adding the 1.0% yield extension to the percent extension between the jaws gripping the strand at rupture, as specified by ASTM A416 (ASTM 2018a). The strands had an average yield strength f¯py of 1,690 MPa (246 ksi), an average ultimate tensile strength f¯pu of 1,932 MPa (281 ksi), and an average rupture strain ε¯pu of 0.0674. The coefficient of variation (COV) was 0.9% for f¯py and f¯pu, and 5.1% for ε¯pu. The strands modulus of elasticity Ep was assumed to be 196.5 GPa (28,500 ksi). Based on these results, the strands conform to the mechanical property thresholds of ASTM A416 for Grade 1860 MPa (270 ksi) strands.

Test Results, Analysis, and Observations

The applied moment versus deflection curves at the east and west point loads are shown in Fig. 6(a), with the unloading and reloading portions at the transitions between the three test stages removed for clarity. The presented applied moment M is equal to the east and west load cells readings multiplied by the flexural span of 8.69 m (28.5 ft) and does not include the initial moment at the midspan Mini induced by the self-weight of the girder and the loading apparatus (e.g., spreader beam, transfer plates, load cells). The initial moment was equal to 467 kN-m (345 kip-ft) and was calculated based on the difference in the load cell readings before and after the girder and loading apparatus were installed. The east and west reaction point deflections were calculated as the vertical displacement near the east and west load cells (WP2 and WP3) with respect to a reference line connecting the east and west supports. The data from the east and west load cells correlated well, as is shown in Fig. 6(a). The average curve in Fig. 6(a) represents the average applied moment versus the average deflection behavior of the constant moment region and was obtained by averaging the load and displacement results at the east and west reaction points throughout the test.
Fig. 6. Relationships between applied moment as a function of (a) the deflection of the constant moment region; and (b) the change in the longitudinal strain at extreme compression and tension fibers in each of the five instrumented regions (full unloading and reloading cycles omitted for clarity).
The average midspan moment-deflection response of the girder was linearly elastic until softening started to occur at an applied moment of approximately 3,655 kN-m (2,695 kip-ft) and a deflection of 65.9 mm (2.60 in.), indicating the initiation of flexural cracking in the girder. As the load continued to increase, the girder continued to sustain inelastic damage and exhibited a gradual decrease in the flexural stiffness. This behavior was likely due to the development of tight and closely spaced cracks within the tensile zone at the bottom half of the section. These cracks were periodically investigated by halting the loading protocol during the first two test stages, up to an applied moment of 6,920 kN-m (5,104 kip-ft), and by spraying the girder’s bottom flange and web within the constant moment region with an evaporative liquid (denatured alcohol). The liquid briefly penetrates the cracks while evaporating from the girder’s surface, temporarily making the cracks visible. However, none of the flexural cracks were visible to the naked eye during the first two test stages, indicating extremely tight cracks that were efficiently controlled by the fiber and strand reinforcements. The detection of cracks was not investigated during the third test stage due to safety concerns.
The flexural capacity of the girder was achieved at an applied moment of 7,271 kN-m (5,362 kip-ft) and a midspan deflection of 234 mm (9.21 in.). As the capacity was reached, a localized crack appeared at the midspan, initiating at the extreme bottom fiber and propagating swiftly up through the bottom half of the web. The localization of the crack resulted in an abrupt decrease of approximately 13.2% in the applied moment as shown in Fig. 6(a). A photo of the crack pattern at the midspan immediately after localization is shown in Fig. 7(a) (crack pattern shown for this reading is indicated by a star in Fig. 6). After localization, the girder began to hinge about the cross section at the localized crack. This behavior is unique to strain-hardening fiber-reinforced concrete beams, wherein: (1) the significant increase in curvature at the localized crack will contribute to a major portion of the overall deflection of the beam; and (2) the curvature at the nonlocalized cracks will be controlled by the unloading and reloading stiffness of these sections where the UHPC is exhibiting prelocalization behavior. The large deformations at the localized crack cross section locally strained the bottom strands until they ruptured, separating the girder into two pieces only connected by the top two strands. A photo of the girder after strand rupture is shown in Fig. 7(b). After the localization-induced applied moment reduction, the applied moment marginally increased with increasing displacement until the strands ruptured at an applied moment of 6,467 kN-m (4,769 kip-ft) and midspan deflection of 287 mm (11.3 in.).
Fig. 7. Photos of the cracking pattern of the girder at midspan showing the localized crack (a) immediately after the drop in peak load; and (b) after the rupture of the bottom strands.
The pair of top and bottom LVDTs that were installed in each of the five instrumented regions of the girders (Fig. 4) were used to create the average linear strain profiles in these regions throughout the duration of the test (assuming plane sections before bending remain plane after bending). Each strain profile was then utilized to determine the change in concrete strains at extreme top (compression) and bottom (tension) fibers within the respective region as a function of the average applied moment at the midspan, as shown in Fig. 6(b). The relationships in Fig. 6(b) confirm that, after the localized crack initiated within the constant moment region (Region 3) at the peak moment, the deformation of the cross section at the localized crack increased, as evidenced by the strain measurements at the top and bottom of the section, while the neighboring cracks closed as the applied load decreased. Note that the longitudinal deformation after the formation of a localized crack should be discussed in terms of crack opening rather than strain; the postpeak strain values of Region 3 in Fig. 6(b) reflect the observed longitudinal extensions within the gauge length.
The stress in the prestressing strands at the beginning of the flexural test, including all losses between strand release and time of test, was estimated by utilizing the concrete strains measured by four vibrating wire gauges (VWGs) cast into the midspan cross section. Two gauges were embedded between the top strands and two at the centroid of the bottom strands. The initial concrete strain profile was then constructed by comparing the concrete strains at the time of test to the reference strains taken prior to the detensioning of the strands. The difference in concrete strains at the level of each layer of strands was then determined assuming a perfect bond between the strands and the surrounding concrete. When these strains are multiplied by the modulus of elasticity of the strands, Ep=196.5  GPa (28,500 ksi), the difference in the stresses in the strands between jacking and the time of test can be estimated. The effective prestress in the first (top), second, and third (bottom) layers of strands at the midspan at the start of the test was calculated to be 1,336  MPa (193.7  ksi), 1,157  MPa (167.8  ksi), and 1,144  MPa (166.0  ksi), corresponding to strains of εp1,ini=-0.00680, εp2,ini=0.00589, and εp3,ini=0.00582, respectively (tension taken as a negative value).
The initial stress profile in the girder at the beginning of the test was obtained from an elastic analysis of the cross section at the midspan and using a concrete modulus of elasticity of 43.6 GPa (6,317 ksi), taken as the average of the two elastic modulus values reported for each batch (Table 1). The initial stresses at the midspan were σT,ini=0.94  MPa (0.136 ksi) at the extreme top fiber (compression taken as a positive value) and σB,ini=29.7  MPa (4.31 ksi) at the extreme bottom fiber, corresponding to strains of εT,ini=0.000022 and εB,ini=0.000682, respectively.
A summary of the test results at the midspan, including the effect of the prestressing force and considering the self-weight of the girder and loading apparatus, is calculated at the five points in flexural behavior, namely, the initiation of cracks; service stress limit of prestressing steel taken when the stress in the bottom layer of strands is equal to 80% of the strand yield stress (AASHTO 2020), i.e., fp3=0.80f¯py; yielding of the bottom layer of strands, i.e., fp3=f¯py; crack localization; and strand rupture as presented in Table 2. In Table 2, Mexp is total moment corresponding to applied moment M plus the self-weight moment Mini, as described in Eq. (1); εT and εB are the total concrete strains in the top and bottom layers calculated according to Eqs. (2) and (3), in which ΔεT and ΔεB are the change in strains captured in extreme top and bottom fibers during the test within Region 3; εp3 is the strain in the bottom layer of strands calculated according to Eq. (4), in which εp3,ini=0.00582 is the initial strain in the bottom layer of strands and Δεcp3 is the change in strain in the concrete at the level of the bottom layer of strands during the test within Region 3; X is the neutral axis depth calculated from the strain profile formed by εT and εB; and ψ=εT/X is the sectional curvature at the midspan. The 80% of yield and yield moment values reported in Table 2 and Fig. 6 were determined when εp3 reached a strain equal to -0.00705 (corresponding to 0.80fpy) and the yielding strain of the strands, εpy=0.01, respectively
Mexp=M+Mini
(1)
εT=εT,ini+ΔεT
(2)
εB=εB,ini+ΔεB
(3)
εp3=εp3,ini+Δεcp3
(4)
Table 2. Summary of test results at midspan considering the effect of the prestressing force and considering the self-weight of the girder and loading apparatus
ParameterFirst crack80% of yieldbYieldcLocalizationStrand rupture
Mexp kN-m (kip-ft)4,122 (3,040)4,884 (3,602)6,983 (5,150)7,775 (5,734)6,970 (5,141)
εTa0.000990.001230.002380.002990.00482d,e
εBa0.000360.000720.003760.006180.03143d,e
εpb3a0.006720.007050.010000.012390.03611d,e
X mm (in.)660 (26.0)568 (23.4)349 (13.8)293 (11.5)120 (4.72)e
ψ1/mm (1/in.)1.50×106 (3.80×105)2.16×106 (5.49×105)6.82×106 (1.73×104)1.02×105 (2.58×104)4.02×105 (1.02×103)e
a
Compression strains are taken as positive values and tension strains are taken as negative values.
b
Calculated when the stress in the bottom layer of strands is equal to the service stress limit of prestressing steel, i.e., fp3=0.80f¯py, as defined in Table 5.9.2.2-1 of AASHTO LRFD BDS (AASHTO 2020).
c
Calculated when the stress in the last layer of strands is equal to the yielding stress limit of prestressing steel, i.e., fp3=f¯py.
d
Postpeak deformation accumulated in the localized crack; strain values reflect the longitudinal extensions within the gauge length of the LVDTs.
e
Calculations assume strain compatibility is maintained after localization.
Finally, the slippage of selected strands during the test was monitored with three LVDTs mounted on three of the bottom-layer strands located within one half of the section on the west end of the girder. As expected with a beam of this length, these measurements demonstrated that no strand slippage occurred during the test.

Description of Proposed Flexural Design Method for UHPC Beams

The proposed flexural design framework described in this section is relevant to the design of beams made with UHPC and reinforced with conventional steel and/or prestressing strands. The framework is founded on structural mechanics principles commonly used in conventional reinforced concrete design where the nominal flexural capacity at each cross section of the beam is calculated by satisfying the conditions of equilibrium and strain compatibility. Because the mechanical behavior of UHPC is different from that of conventional concrete, new constitutive relationships are proposed and then utilized within the strain compatibility approach to predict the complete nominal moment–sectional curvature (Mn-ψ) diagram and the flexural failure modes of UHPC beams. In this framework, it is assumed that (1) sections plane before bending remain plane after bending; (2) the member is reinforced with conventional steel and/or prestressing strands within the flexural tension side; (3) a perfect bond exists between the reinforcement and UHPC, i.e., the change in strain in the bonded reinforcement and/or prestressing strands is equal to the strain in the surrounding UHPC; (4) the stresses in the reinforcing steel or prestressing strands are derived from the uniaxial stress-strain response indicative of their tensile behavior; (5) the composition of UHPC mixture and its mechanical properties meet specified threshold values; and (6) the stresses in the UHPC are inferred from appropriate uniaxial stress-strain models. The specified threshold values and the uniaxial stress-strain models for UHPC are discussed in the following section.

UHPC-Class Materials and Mechanical Models

A key starting point for the proposed flexural design framework is the delineation of the scope of the considered UHPC-class materials and the definition of the fundamental parameters. UHPC is generally defined as portland cement composite made of an optimized gradation of granular constituents, a water-to-cementitious materials ratio less than 0.25, and a high percentage of discontinuous internal steel fiber reinforcement. The specific mechanical properties include a minimum compressive strength of 124 MPa (18.0 Ksi), a minimum cracking strength of 5.0 MPa (0.73 Ksi), and the ability to sustain the cracking strength through a minimum localization strain of at least 0.0025.
The compressive properties, namely, the modulus of elasticity Ec, the compressive strength fc, and the strain at compressive strength εcu, are obtained from cylindrical specimens tested according to ASTM C39 (ASTM 2018b) and ASTM C469 (ASTM 2014) with provisions specific to UHPC described in ASTM C1856 (ASTM 2017). The uniaxial elastic-perfectly plastic model proposed by El-Helou et al. (2022) is adopted in this framework to determine the stress in the UHPC as a function of the compression strain in the section, as shown in Fig. 8(a). This model mimics the experimental uniaxial stress-strain response in the elastic region up to a reduced compressive stress of αfc, where α is a reduction factor on compressive strength reflecting the linearity limit of the compressive stress-strain response. After this stress level, the model sustains the reduced compressive resistance until the strain at the material’s compressive strength εcu is reached. El-Helou et al. (2022) suggested a reduction factor α value of 0.85, based on the experimental results of five commercially available UHPC products (Graybeal and Stone 2012; Haber et al. 2018).
Fig. 8. Idealized stress-strain relationships for UHPC subjected to (a) uniaxial compression and uniaxial tension with (b) postcracking stress plateau; and (c) strain hardening with continuous increase in postcracking stress.
The tension parameters, namely, the effective cracking strength ft,cr, the localization stress ft,loc, and the localization strain εt,loc, are obtained from prismatic specimens tested in direct tension according to the method developed by Graybeal and Baby (2013). This method delivers the uniaxial stress-strain responses of the tested specimens, facilitating the identification of critical tension parameters. The effective cracking strength ft,cr, is taken as the intercept of a line having a slope equal to the elastic modulus and a strain offset of 0.02%. The localization stress ft,loc and strain εt,loc are taken as the stress and strain of the data point where the stress begins to continuously decrease with increasing strain. As proposed by El-Helou et al. (2022), the constitutive law for UHPC in tension idealizes the experimental curve into an elastic-perfectly plastic relationship, as shown in Fig. 8(b), for material exhibiting a stress plateau after cracking, or a bilinear relationship, as shown in Fig. 8(c), for material exhibiting a continuous increase in stress after cracking and when the localization stress ft,loc is at least 20% greater than the effective cracking stress ft,cr. El-Helou et al. (2022) proposed that a reduction factor γ be applied on the tensile stress parameters (ft,cr and ft,loc) to account for variability in the material behaviors; the value of the reduction factor γ is not to exceed 0.85.
For the conventional steel reinforcement, an elastic-perfectly plastic stress-strain relationship is adopted, as shown in Fig. 9(a), in which Es is the modulus of elasticity of conventional steel; fsy and εsy are the yielding stress and strain, respectively; and εsu is the strain at rupture of the bars. The idealized stress-strain relationships for prestressing strands can be obtained from existing nonlinear models, such as the power formula based on the work of Skogman et al. (1988) shown in Fig. 8(b). For instance, the stress in the prestressing steel fps for 1,860 MPa (270 ksi) seven-wire low-relaxation strands can be computed as a function of the steel strain εps according to Eq. (5). To obtain fps in kilopounds per square inch in Eq. (5), replace 6,116 MPa with 887 ksi, 190,385 MPa with 27,613 ksi, and 1,860 MPa with 270 ksi (PCI 2014)
fps=εps[6,116+190,385(1+(112.4εps)7.36)17.36]1,860  MPa
(5)
Fig. 9. Idealized uniaxial stress-strain relationships for (a) conventional steel reinforcing bars; and (b) prestressing strands.

Flexural Behavior of UHPC Beams

The flexural behavior of UHPC beams can be obtained analytically by employing a strain compatibility approach and utilizing the constitutive models for UHPC and steel reinforcement depicted in Figs. 8 and 9, respectively. For flexural members fully utilizing the UHPC tensile resistance, the nominal moment–sectional curvature (Mnψ) diagram can be idealized in four key points, as illustrated in Fig. 10, for a UHPC member with conventional steel reinforcement. The sectional curvature is defined as the ratio of the strain in the extreme compression layer εc divided by the depth of the neutral axis X: ψ=εc/X. The first key point in the behavior (ψcr, Mcr) represents the initiation of the first flexural tensile crack, occurring when the strain in the extreme tension layer reaches the cracking strain limit of UHPC, εt,cr=ft,cr/Ec. At this point, the compression and tensile stresses remain within the elastic region of their respective constitutive models (Fig. 8). The second point in the moment-curvature diagram (ψy, My) coincides with the yielding of the extreme tension layer of the reinforcing steel, occurring when the steel strain is equal to the yielding strain of conventional steel εsy or prestressing strands εpy. Note that in Fig. 10, an intermediate point (ψsl, Msl) in behavior is shown corresponding to the chosen baseline sectional curvature and moment for ductility considerations as discussed later in the manuscript. The third key point (ψL, ML) indicates the flexural nominal capacity of the section (Mn=ML) and coincides with the crack localization, where the tensile capacity of UHPC is fully utilized and the strain in the extreme tension layer is equal to the localization strain εt,loc. The cross-sectional strain and stress diagrams at the flexural crack localization point for a UHPC material exhibiting a stress plateau after cracking [Fig. 8(b)] are shown in Fig. 11, where the compression stresses remain elastic with the strain value in the extreme compression layer εc, less than εcp=αfc/Ec [Fig. 11(a)], or become plastic when εcp<εcεcu [Fig. 10(b)]. In Fig. 10, it is assumed that the localization strain of UHPC εt,loc is greater than the yielding strain of the reinforcement εsy, and thus the yielding point (ψy, My) occurs at a smaller curvature and moment values than those defining the localization point (ψL, ML). For prestressed members, the yielding point occurs before the localization point if the change in the prestressing steel strain between the effective prestress and yield (εpyεpe) is greater than the UHPC localization strain εt,loc. After crack localization (εt>εt,loc), the UHPC can no longer contribute to the tension resistance of the member, leading to a reduction in the flexural capacity. The last point in the flexural behavior shown in Fig. 10 (ψc, Mc) coincides with the compression failure of UHPC, occurring when the strain in the extreme compression layer reaches the ultimate compressive strain of UHPC εcu. The cross-sectional strain and stress diagrams at compression failure and after the localization of cracks (εt>εt,loc) are depicted in Fig. 12 for a UHPC material exhibiting a stress plateau after cracking [Fig. 8(b)]. The shape of the postlocalization moment-curvature relationship is nonlinear (dashed line in Fig. 10) and depends on the UHPC material properties and the design parameters of the section. Intermediate points between localization and compression failure need to be computed if the exact shape of the postlocalization curve is desired.
Fig. 10. Idealized nominal moment versus sectional curvature (Mnψ) diagram of UHPC member with conventional steel reinforcement.
Fig. 11. Cross-sectional strain and stress conditions for UHPC members in flexure at the onset of crack localization with (a) elastic stresses in compression; and (b) plastic stresses in compression (shown for a UHPC exhibiting a stress plateau after cracking in tension).
Fig. 12. Cross-sectional strain and stresses conditions for UHPC members in flexure at compression failure and after the localization of cracks (shown for a UHPC exhibiting a stress plateau after cracking in tension).
In the determination of the moment-curvature diagram in Fig. 10, it is assumed that the strain in the steel reinforcement εs remains less than the strain at rupture of the bars εsu when the compressed UHPC crushes. However, due to the high compressive strength of UHPC materials, the rupture of the steel reinforcement (conventional steel or prestressing strands) can occur before concrete crushing (i.e., when the strain in the extreme compression layer is less than εcu). In an effort to engage the full compressive strength of UHPC in a flexural design, Shao and Billington (2019) demonstrated that the postlocalization moment capacity of UHPC beams can be carried through high-tensile strains until compression crushing (McML) when the beam is heavily reinforced with conventional steel reinforcement and when the strain-hardening characteristics of the steel are both engaged and substantial. While such behavior is desirable because it significantly increases the ductility of UHPC beams, to date the approach has only received limited experimental investigation. When the tensile strain in the UHPC exceeds the localization strain at a crack (εt>εt,loc), the loss of the UHPC fiber-bridging resistance mechanism results in a change in the internal force-resisting mechanism (i.e., increased curvature, increased strains, and movement of the neutral axis). If the increased tensile strain results in increased tensile resistance due to the engagement of additional discrete reinforcement force-resisting mechanisms, such as the strain-hardening of steel or the further engagement of unbonded reinforcements, then a new stable moment-resisting state may be realized. For this reason, postlocalization analysis should take into consideration the width of the localized crack, the local straining of the reinforcement crossing the crack, and the resistance provided by the reinforcement. In the current proposal, the postlocalization flexural capacity is not recommended for use in design because more research is needed to quantify the relationship between the rupture strain of the bars after UHPC crack localization, the opening of the localized crack, and the development length of the reinforcement in the vicinity of the localized crack. The development length of the reinforcement is a critical consideration as it significantly influences the length of the bar or strand over which the increased tensile strain demand will be resisted.
For flexural members not fully utilizing the UHPC tensile resistance, the ultimate capacity occurs at compression crushing (Mn=Mc), when the UHPC in the extreme compression layer reaches its ultimate compressive strain limit (εc=εcu) while the strain in the extreme tensile fiber is less than the localization strain (εt<εt,loc). In these situations, the beam design does not take full advantage of the UHPC tensile performance characteristics, prompting a compression failure at reduced tensile strain values, such as could occur in sections having a very high reinforcement ratio and/or sections subjected to high axial compressive loads. The nominal flexural moment–sectional curvature (Mnψ) for this failure mode can be defined by straight lines, with either (1) three key points coinciding with the initiation of first flexural tensile crack (ψcr, Mcr), yielding of the extreme tension layer of the reinforcing steel (ψy, My), and the compression crushing of UHPC (ψc, Mc) when the steel strain at crushing is greater than the yield strain; or (2) two key points coinciding with the initiation of cracks (ψcr, Mcr) and compression crushing (ψc, Mc) when the steel strain at crushing is less than or equal to the yield strain.

Proposed Design Methodology

Based on the experimentally observed behaviors and analytical results, a flexural design framework for UHPC beams reinforced with conventional steel or prestressing strands is proposed in this section. The design framework is similar to portions of existing UHPC structural design guidance detailed in various international documents [NF P18-710 (AFNOR 2016a); SIA 2052 (SIA 2016); CSA S6:19 (CSA 2019)] and is tailored to generally parallel existing bridge design specifications in the United States (AASHTO 2020) while accounting for UHPC mechanical performance.
The first step in the structural design is the characterization of the material-level properties and the establishment of constitutive laws that describe the mechanical response when subjected to structural loads. For UHPC in compression, the parabolic stress-strain relationship typically used for conventional concretes is replaced with the idealized elastic-perfectly plastic model in Fig. 8(a). The compression model requires the modulus of elasticity Ec, compressive strength fc, and strain at compressive strength εcu as input parameters, which can be experimentally obtained from compression tests performed according to ASTM C1856 (ASTM 2017). In lieu of experimental testing, the elastic modulus can be approximated by Eq. (6) and εcu can be taken as the greater of εcp=αfc/Ec and 0.0035, as proposed by El-Helou et al. (2022). The reduction factor α can be taken as 0.85
Ec=9,100fc0.33  (MPa)Ec=2,500fc0.33(ksi)
(6)
In tension, UHPC offers a postcracking resistance that is sustained through large tensile strains (εt>0.0025), delivering a beneficial contribution to the flexural capacity of a beam. The tensile constitutive models for UHPC can be idealized with the elastic-perfectly plastic or bilinear (when ft,loc1.20ft,cr) relationships shown in Figs. 8(b) and 7(c), respectively. The tensile parameters, namely, the effective cracking stress ft,cr and the localization stress ft,loc and strain εt,loc, can be obtained by executing direct tension tests proposed by Graybeal and Baby (2013).

Service and Ultimate Limit States

The flexural analysis can be performed using a strain compatibility approach where stresses over strained cross sections are calculated and then summed to satisfy equilibrium requirements. In a situation where the cracking of the girder is not allowed at service, the linearly elastic analysis procedures typically used with conventional concrete design can be employed, with a tensile stress limit taken as a fraction of the effective cracking stress, e.g., ft,lim=0.85ft,cr. When a minimal amount of cracking is permissible at service, a strain compatibility analysis can be performed in which the strains in the tensile zone are limited to a fraction of the localization strain, e.g., εt,lim=0.30εt,loc, and the stresses in the reinforcing steel are lower than a fraction of the yielding stress, e.g., fs,lim<0.80fsy and fp,lim<0.80fpy. In both service limit state cases, the compression stress limits outlined in AASHTO LRFD BDS (AASHTO 2020) for conventional concrete can also be used for UHPC. For the ultimate limit state, the stress-strain constitutive models in Fig. 8 are used, allowing the construction of the full moment-curvature diagram. In this case, the linear strain distribution along the height of the section is bounded by the UHPC crushing strain limit εcu, the localization strain εt,loc, and the rupture strain limit of conventional steel εsu or prestressing strands εpu. The nominal moment capacity is defined at either (1) crack localization (Mn=ML), when the strain in the extreme tensile layer εt is equal to εt,loc, which can occur when the compression stress is elastic (εcεcp) or plastic (εcp<εcεcu), as shown in Fig. 11; or (2) compression failure (Mn=Mc) when the strain in the extreme compression layer εc is equal to εcu and the strain in the extreme tensile layer εt is less than εt,loc.

Consideration of Curvature Ductility

In cases where the flexural failure is governed by the localization of the cracks (Mn=ML), it is important that members exhibit adequate ductility to ensure that large deformations occur before failure of the member. An approach based on sectional curvature ductility can be adopted in which a curvature ductility ratio μ is defined and compared to a specified minimum curvature ductility ratio, e.g., μmin=3.0. The curvature ductility ratio can be taken as the ratio of the sectional curvature at localization ψL and a baseline sectional curvature, ψsl: μ=ψL/ψsl. It is proposed that the baseline sectional curvature ψsl be calculated when the stress in the extreme tension layer of reinforcement is equal to 80% of the yielding stress of the steel reinforcement, i.e., 0.80fpy or 0.80fsy. For prestressed members, this limit corresponds to maximum stress that can be attained in the prestressing steel at the service limit state specified in AASHTO LRFD BDS (AASHTO 2020). When members have a curvature ductility ratio less than μmin, a reduced resistance factor on the nominal moment Mn is recommended, in recognition of the nonductile failure mechanism akin to compression-controlled failure. The proposed resistance factors are discussed later in the manuscript.
Limited research on reinforced UHPC members has shown that the localization strain value is increased when the cracking resistance provided by the fibers is supplemented by conventional steel rebars or prestressing strands (Oesterlee 2010; Baby et al. 2014). For instance, the Swiss recommendation allows the use of a localization strain in the tensile mechanical model that is twice the value obtained from material testing (SIA 2016). In the current study, the measured strain at localization of the flexural crack (Table 2) was approximately 24% greater than the localization strain obtained from a direct tension test (Batch A in Table 1). The implementation of a localization strain value higher than the one obtained from tension testing in flexural design would allow for the use of a larger tensile strain, thus increasing the likelihood of satisfying the ductility requirement for many UHPC beams. However, more research is needed before broad implementation of higher localization strains can be recommended because the percent increase in the usable strain depends on the uniaxial localization strain value obtained from material tests and on the amount of discrete tensile reinforcement in the flexural element. Finally, as previously discussed, the true ductility of a member may be greater than that analytically predicted using the aforementioned procedure if the opening of a localization crack generates sufficient increased resistance from the tensile reinforcement (e.g., through engagement of strain-hardening behaviors in mild steel reinforcement) that the moment at compression crushing Mc becomes greater than ML. This topic is the subject of future research by the authors.

Consideration of Fiber Orientation

The orientation of the fibers within the tensile zone of beams has a significant effect on the tensile response of UHPC. Fibers tend to preferentially align with the direction of flow of fresh UHPC during casting, with dissimilar tensile stress and strain capacities being exhibited between the flow and nonflow directions (Maya Duque et al. 2016; Doyon-Barbant and Charron 2018; Huang et al. 2018; Walsh et al. 2018; Qiu et al. 2020b). For this reason, it is recommended that a casting method be specified by the designer to ensure that the fiber orientation in the in-place material is similar to the one achieved in the material control specimen. Unfavorable fiber orientation with respect to the flexural tensile strains can reduce the expected cracking stress and localization strain capacities of the UHPC in the structure. In these cases, a reduction factor γ lower than the maximum value of 0.85 is recommended to be applied to the stress and strain parameters of the tensile constitutive law, as shown in Figs. 8(b and c), to ensure a conservative representation of the tensile response in the structure. Determining the value of γ to address the fiber orientation effects may, in some instances, require prototype testing as recommended by the French design standard NF P18-710 (AFNOR 2016a).

Factored Flexural Resistance

At present, there is insufficient data to allow for the calibration of the resistance factor for UHPC beams. However, the LRFD framework requires that a factor be applied to the resistance calculation. Thus, it is proposed that the factored flexural resistance Mr of a UHPC beam can be taken in accordance with the relationship shown in Eq. (7), with ϕf being the resistance factor for flexural moment
Mr=ϕfMn
(7)
For members dominated by compression crushing (i.e., Mn=Mc with εc=εcu and εt<εt,loc), the resistance factor can be taken equal to 0.75. This value was chosen to correspond to the resistance factor specified for compression-controlled sections in AASHTO LRFD BDS (AASHTO 2020). For members dominated by crack localization (i.e., Mn=ML with εcεcu and εt=εt,loc), the resistance factor can be taken as 0.90 for sections satisfying the ductility requirement (μμmin) and 0.75 for nonductile sections with a sectional ductility ratio less than or equal to 1.0. For prestressed and non-prestressed members falling in between the two extremes, a linear variation of the resistance factor is proposed in Eq. (8). This transition zone is proposed akin of the transitional zone between compression-controlled and tension-controlled sections for conventional concrete members in AASHTO LRFD BDS (AASHTO 2020)
ϕf=0.75+0.15(μ1.0)μmin1.0;0.75ϕf0.90
(8)

Validation of the Proposed Flexural Model for UHPC Beams

The capability of the model to predict the flexural behavior of prestressed and non-prestressed UHPC beams was evaluated by comparing the model predictions to the experimental data from 45 flexural tests performed by the authors or found in the literature.

Prestressed Girder Test Presented in this Paper

The predicted moment-curvature behavior of the prestressed girder tested by the authors, and previously discussed in this paper, was obtained by discretizing its cross section into 2.5-mm-(0.1-in.)-thick horizontal layers through the height and assuming constant strains within each layer. The initial strains in each layer at the beginning of the test were determined from the UHPC strain measured by the VWGs embedded at the midspan and following the analytical considerations described previously. That is, the initial strain in the first (top), second, and third (bottom) layer of strands is equal to εp1,ini=0.00680, εp2,ini=0.00589, and εp3,ini=0.00582, respectively (tension taken as a negative value). The initial strain in each layer of UHPC was calculated from the initial strain profile defined by the top strain, εT,ini=0.000022, and bottom strain, εB,ini=0.000682 (compression taken as a positive value). The total strain in the UHPC section was obtained by superposing the initial strain profile with an assumed strain profile, i.e., εi=εi,ini+Δεi, in which εi, εi,ini, and Δεi are the total, initial, and assumed strains in the UHPC layer i, as shown in Fig. 13. Similarly, the total strains in the strand layers, εp1, εp2, and εp3, were calculated by superposing the initial steel strains, εp1,ini, εp2,ini, and εp3,ini, with the change in strain in the UHPC at the level of each steel layer, Δεcp1, Δεcp2, and Δεcp3, respectively. The stresses in each layer were calculated by employing the mechanical models shown in Figs. 8(a and b) and 9(b). The stresses were then converted into compression and tension forces, which were summed to check force equilibrium in the section. This process was repeated by assuming different strain profiles until equilibrium was satisfied, after which the flexural moment was computed. The procedure was repeated by increasing the values of assumed strains until the full moment–sectional curvature diagram was derived.
Fig. 13. Determination of the total strains in the UHPC as a superposition of the initial strains (at the beginning of the test) and change in strains due to the applied moment.
In calculating the UHPC stresses, the compression and tension parameters were informed from the average results of the material tests performed on companion specimens presented in Table 1. The UHPC modulus of elasticity was taken as the average of the results obtained from the two UHPC batches with Ec=43.6  GPa (6,317 Ksi). The compressive strength and ultimate compression strain were taken from the results of Batch B, which filled the top half of the girder: fc=161  MPa (23.3 ksi) and εcu=0.00401. The tensile parameters were taken from the results of Batch A, which filled the bottom half of the girder, and without a reduction factor (i.e., γ=1.00): γft,cr=γft,loc=9.3  MPa (1.35  ksi) [elastic-perfectly plastic model in Fig. 8(b)] and εt,loc=0.00497. The tensile design model is compared to the uniaxial tensile test results in Fig. 5(a). The steel stress-strain model relationship for the 1,860 MPa (270 ksi) low-relaxation strands was determined according to Eq. (5) and depicted atop the experimental data from the strand pullout tests in Fig. 5(b). For strand stress values greater than 1,860 MPa (270 ksi) or strain values greater than 0.0272, a linear interpolation was implemented up to a rupture stress fpu of 1,932 MPa (281 ksi) occurring at a strain εpu of 0.0674, as shown in Fig. 5(b). Table 3 presents a summary of the strain compatibility analysis results at the five points in flexural behavior, namely, the initiation of cracks; strand service stress limit, i.e., 0.80fpy, reached in the bottom layer of strands; yielding of the bottom layer of strands; crack localization; and compression crushing. These results represent the nominal design capacities of the girder and can be directly compared to the experimental data summarized in Table 2. The predicted-over-experimental flexural moment ratio, Mn/Mexp, is 0.97, 1.09, 0.99, and 0.92 at first crack, 80% of yield, yield, and localization, demonstrating the capability of the proposed methodology to predict the flexural capacity of prestressed UHPC beams. The predicted nominal design curve is compared to the experimental results in Fig. 14, plotted as the relationship between the applied moment and the change in curvature during the test and simulation. The simulated design and experimental moments in Fig. 14 are taken as MnMini and MexpMini, respectively, where Mini=467  kNm (345 kip-ft) is the initial moment in the beam at the beginning of the test. The change in curvature Δψ is taken as ΔεT/ΔX, where ΔX is the depth of the neutral axis of the intermediate strain diagram shown in Fig. 13, a value calculated in the strain compatibility analysis and measured during the test. Note that the simulated design curve was obtained from 110 calculation points, which is greater than the minimum four key points proposed in the design methodology.
Table 3. Summary of the predicted results at midspan computed according to the proposed flexural framework
ParameterFirst crack80% of yieldbYieldcLocalizationCrushingd
Mn kN-m (kip-ft)4,015 (2,961)5,306 (3,913)6,921 (5,105)7,132 (5,260)6,636 (4,894)
εTa0.000990.001390.002260.002450.00401
εBa0.000210.000730.003910.004970.02262
εpb3a0.006610.007080.010000.011020.02759
X mm (in.)741 (29.2)590 (23.2)330 (13.0)297 (11.7)136 (5.3)
ψ1/mm (1/in.)1.34×106 (3.40×105)2.35×106 (5.97×105)6.85×106 (1.74×104)8.23×106 (2.09×104)2.96×105 (7.51×104)
Mn/Mexp0.971.090.990.920.95
a
Compression strains are taken as positive values and tension strains are taken as negative values.
b
Calculated when the stress in the bottom layer of strands is equal to the service stress limit of prestressing steel, i.e., fp3=0.80f¯py, as defined in Table 5.9.2.2-1 of AASHTO LRFD BDS (AASHTO 2020).
c
Calculated when the stress in the last layer of strands is equal to the yielding stress limit of prestressing steel, i.e., fp3=f¯py.
d
Calculations assume strain compatibility is maintained after localization.
Fig. 14. Relationship between the experimental applied moment and the change in sectional curvature at midspan, comparing (a) experimental and design responses; and (b) experimental and fitted experimental responses.
The experimental and predicted design responses in Fig. 14(a) are in good agreement, demonstrating the capability of the proposed method to predict the full moment–sectional curvature response of prestressed beams. The results also confirm that the shape of the moment–sectional curvature diagram can be idealized by four key points, as illustrated in Fig. 10. Because the compression strain limit observed during the test was approximately 20% larger than the ultimate compression strain specified in the simulations, the predicted failure mode of the girder was compression crushing instead of strand rupture, as observed in the test [Table 3 and Fig. 14(a)]. This result suggests that the UHPC in a structural element can undergo compression strains exceeding the strain at peak stress, as obtained from a uniaxial compression test, likely due to the confinement effects within the larger beam and strain gradient effect within a beam cross section with large curvature. Similarly, the localization strain of UHPC observed during the test (εB=0.00618 in Table 2 at localization point) was approximately 24% greater than the localization strain obtained from direct tension testing [εB=ε¯t,loc=0.00497 in Table 1 (Batch A) and Table 3 at localization point], indicating that the primary tensile reinforcement and strain gradient effects may enhance the flexural tensile strain capacity of the UHPC.
The design curvature ductility ratio of the girder, μsim=3.50, can be calculated by dividing the sectional curvature at localization over the sectional curvature at 80% of yield reported in Table 3. Because the ductility ratio is greater than the specified ratio μmin of 3.0, the flexural resistance factor for this girder is equal to 0.90. Note that the experimental ductility ratio obtained from the results reported in Table 2 is greater than μsim with μexp=4.72 because the cracks in the girder localized at a higher localization strain than the one used in the simulations. The factored moment-curvature design curve is compared to the nominal and experimental curves in Fig. 14(a).
To illustrate the key differences between the behavior of prestressed conventional concrete beams and similar UHPC beams, a strain compatibility analysis of the same girder was performed ignoring the tensile resistance of UHPC, i.e., ft,cr=ft,loc=0, with the results plotted in Fig. 14(a). While this analysis resulted in a nominal moment capacity (Mn=Mc) lower than the experimental capacity, it both incorrectly predicted a concrete crushing failure mode at the top of the section, rather than localization of cracks at the bottom, and also significantly overestimated the curvature ductility ratio μnt,sim compared to the experimental ductility ratio, i.e., μnt,sim=2.67μexp. This comparison demonstrates that conducting a capacity-based design that ignores the UHPC tensile resistance may inadvertently result in a structure that does not demonstrate the desired ductility.
Finally, if the localization strain measured during the girder test, εt,loc=0.00636, is used in the model, the simulated nominal design curve is identical to the prelocalization experimental moment-curvature curve, as shown in Fig. 14(b), when the bilinear constitutive law in Fig. 8(c) is employed with the following calibrated input values for the cracking and localization stresses: ft,cr=6.21  MPa (0.90 ksi) and ft,loc=15.5  MPa (2.25  ksi). In this analysis, the estimated value of the cracking stress in the girder was approximately 33% lower than the one obtained from direct tension testing, while the localization stress was 49% higher than the localization stress value obtained from tension testing (Batch A in Table 1).

Prestressed and Non-Prestressed Beams Tested by Other Researchers

This section explores the experimental results from tests of two additional prestressed girders and 42 non-prestressed reinforced beams found in the literature and compares them to model predictions. The beams had varying cross-sectional shapes and geometries (rectangular shapes for non-prestressed, and I and pi shapes for prestressed). The height of the non-prestressed rectangular beams ranged between 220 mm (8.7 in.) and 381 mm (15 in.), while the height of the prestressed I- and pi-shaped beams were 838 mm (33 in.) and 914 mm (36 in.). The beams were made with different types of UHPC products with varying amounts of fiber volume content (1.0%, 1.5%, and 2%) and prestressing or conventional steel reinforcement (reinforcement ratio ρ ranging between 0% and 4.99%). The fibers varied in shape and geometries, including smooth, twisted, straight, and hooked-end, with diameters and lengths ranging between 0.12 and 0.30 mm (0.0047 and 0.012 in.) and 10.0 and 30.0 mm (0.39 and 1.18 in.), respectively. The reported flexural behaviors of the beams are in line with the flexural behavior of UHPC members described previously in this paper: the maximum flexural moment was reached immediately before a single localized crack—initiated by the pullout of the crack-bridging fibers—occurred, followed by the subsequent rupture of the tensile reinforcement. Therefore, the analytical predictions presented in this section focus on determining the capacity of the beams at the onset of localization, as recommended in the proposed design methodology that is developed in this paper. The predicted results for the beams found in the literature are presented in Table 4. A portion of the input material parameters for the UHPC and reinforcing steel given in Table 4 were obtained from independent material tests that accompanied each beam, i.e., simulations of beams tested by Chen et al. (2018), Yoo et al. (2017), Qiu et al. (2020a), Yoo and Yoon (2015), Graybeal (2006b), and Graybeal (2009). The uniaxial tensile parameters (γft,cr, γft,loc, and εt,loc with γ=1.00) for the UHPC in these beams were obtained from uniaxial testing methods or inverse analysis techniques performed on the results of flexural beam tests. In cases where key material parameters were not reported, they were obtained by fitting the model results to the experimental data from one of the tested beams in a set, then comparing the predicted flexural capacities of the remaining beams within the same set to the experimental results without changing the estimated input parameters, i.e., simulations of beams tested by Yang et al. (2010), Meade and Graybeal (2010), and Hasgul et al. (2018), as indicated in Table 4. The predicted flexural moment capacity Mn of the beams presented in Table 4 is in good agreement with an average predicted-over-experimental moment ratio, Mn/Mexp, of 0.96 with a COV of 19.6% for all beams, providing another confirmation of the capability of the proposed methodology to predict the flexural capacity of prestressed and non-prestressed UHPC beams. The proposed closed-form model inherently depends on the input mechanical properties; more rigorous determination of these properties will result in a more accurate approximation of the flexural capacity.
Table 4. Comparison between experimental and predicted flexural capacities for prestressed and non-prestressed UHPC beams found in the literature
ReferenceOriginal namefc (MPa)E (MPa)εcuft,cr (MPa)ft,loc (MPa)εt,locρ (%)fsy (MPa)Mn (kN-m)Mexp (kN-m)Mn/Mexp
Chen et al. (2018)aB-112645.6f0.0035g9.8h11.9h0.0043h1.0946154.043.31.25
B-212645.6f0.0035g9.8h11.9h0.0043h2.7541778.571.41.10
B-312645.6f0.0035g9.8h11.9h0.0043h3.6045696.290.41.06
B-412645.6f0.0035g9.8h11.9h0.0043h4.99445111.2105.91.05
Yoo et al. (2017)aUH-N19747.80.004411.1i12.2i0.0032i0.00N/A66.572.50.92
UH-0.53%19747.80.004411.1i12.2i0.0032i0.5352391.997.90.94
UH-1.06%19747.80.004411.1i12.2i0.0032i1.06523117.0118.80.99
UH-1.71%19747.80.004411.1i12.2i0.0032i1.71523128.1131.00.98
Qiu et al. (2020a)bB-S65-1613244.00.0035g8.0h9.4h0.0048h0.7341283.865.51.28
B-S81-2012643.10.0035g8.1h9.7h0.0062h1.14460114.0105.01.09
B-S83-2013045.30.0035g8.0h9.6h0.0053h1.14460113.0101.91.11
B-H65-2013945.20.0035g8.0h10.4h0.0067h1.14502123.0133.00.93
Yoo and Yoon (2015)cS13-0.94%21246.70.00457.0i7.0i0.0023i0.9449433.539.30.85
S13-1.50%21246.70.00457.0i7.0i0.0023i1.5050340.255.80.72
S19.5-0.94%21046.90.004810.6i10.6i0.0031i0.9449444.842.01.07
S19.5-1.50%21046.90.004810.6i10.6i0.0031i1.5050353.756.30.95
S30-0.94%21046.80.00469.8i11.0i0.0073i0.9449447.043.21.09
S30-1.50%21046.80.00469.8i11.0i0.0073i1.5050357.456.11.02
T30-0.94%23247.00.00539.5i11.3i0.0036i0.9449447.943.51.10
T30-1.50%23247.00.00539.5i11.3i0.0036i1.5050358.460.30.97
Yang et al. (2010)aNR-1,219746.80.0035g10.7j10.7j0.0070j0.00N/A59.371.00.84
NR12-1,219146.40.0035g10.7j10.7j0.0070j0.60510j85.185.11.00l
R13-1,219246.70.0035g10.7j10.7j0.0070j0.09510j97.9102.00.96
R13C-1,219246.70.0035g10.7j10.7j0.0070j0.90510j97.992.21.06
R14-1,219645.50.0035g10.7j10.7j0.0070j1.20510j110.6116.70.95
R22-1,219146.40.0035g10.7j10.7j0.0070j1.31510j105.5106.40.99
R23-219645.50.0035g10.7j10.7j0.0070j1.96510j128.2131.60.97
Meade and Graybeal (2010)dS1-119452.50.0035g7.1j7.1j0.0045j0.4147697.5102.90.95
S1-219452.50.0035g7.1j7.1j0.0045j0.55476103.6103.51.00l
S1-319452.50.0035g7.1j7.1j0.0045j0.64463113.7121.00.94
S1-419452.50.0035g7.1j7.1j0.004j0.83476122.4123.50.99
S1-519452.50.0035g7.1j7.1j0.0045j1.00448127.5135.50.94
S2-020353.30.0035g10.7j10.7j0.0045j0.00N/A101.590.01.13
S2-120353.30.0035g10.7j10.7j0.0045j0.41476126.0133.80.94l
S2-220353.30.0035g10.7j10.7j0.0045j0.55463135.4120.71.12
S2-320353.30.0035g10.7j10.7j0.0045j0.64476141.2163.40.86
S2-420353.30.0035g10.7j10.7j0.0045j0.83476153.0132.31.16
S2-520353.30.0035g10.7j10.7j0.0045j1.00448156.7171.20.91
Hasgul et al. (2018)eB1-F15749.1f0.0035g6.5j6.5j0.0043j0.9045352.652.61.00l
B2-F16750.1f0.0035g6.5j6.5j0.0043j1.9046380.089.80.89
B3-F15749.1f0.0035g6.5j6.5j0.0043j2.80456100.9111.90.90
B4-F16650.0f0.0035g6.5j6.5j0.0043j4.3046598.9134.40.74
Graybeal (2006b)a80F20052.40.004310.0k11.2k0.0047km18623,4924,8020.73
Graybeal (2009)aP2-70F22156.30.0035g10.0k11.2k0.0047km18623,6164,2480.85

Note: 25.4  mm=1  in.; and 1 lb = 4.45 N.

a
Specimens made of UHPC dosed with 2% straight smooth fibers by volume; fibers had a diameter of 0.2 mm and length of 13 mm.
b
Specimens made of UHPC dosed with 2% fibers by volume; fibers had smooth, twisted, and hooked-end shapes, diameters of 0.12, 0.16, and 0.20 mm, and lengths of 10.0 and 13.0 mm.
c
Specimens made of UHPC dosed with 2% fibers by volume; fibers had smooth and twisted shapes, diameters of 0.20 and 0.30 mm, and lengths of 13.0, 19.5, and 30.0 mm.
d
Specimens made of UHPC dosed with 1% and 2% straight smooth fibers by volume; fibers had a diameter of 0.2 mm and length of 13 mm.
e
Specimens made of UHPC dosed with 1.5% straight smooth fibers by volume; fibers had a diameter of 0.13 mm and length of 13 mm.
f
Values not reported; they were approximated using Eq. (6).
g
Values not reported; they were assumed based on the typical proposed value for the ultimate compressive strain for UHPC.
h
Values reported from independent uniaxial tension tests; note that the test methods utilized were different than the one proposed in this paper.
i
Values reported from inverse analysis of independent flexural beam test results following the recommendations of AFNOR (2016b).
j
Values not reported and were obtained by the calibration of the input parameters using the flexural results of only one beam in each set of specimens.
k
Values taken from uniaxial tension tests of similar UHPC product performed by the same authors: Set F1A-Long in Graybeal and Baby (2013).
l
Specimen flexural test results were used for calibration of a number of input material parameters that were not reported for the specimen set.
m
Prestressed girders with effective prestress of 779 and 1,061 MPa for girders 80F and P2-70F, respectively.

Summary and Conclusions

Aspects of the flexural behavior of UHPC beams are distinct from those commonly associated with conventional concrete beams. This paper presents a flexural design framework based on rational principles with no reliance on empirical formulations. The framework is supported by an experimental investigation of the flexural behavior of a pretensioned bulb-tee bridge girder. Analytical results were compared to experimental data from prestressed and non-prestressed flexure beam tests found in the literature and demonstrate the capability of the proposed design method to predict flexural capacity. Based on these investigations, the following conclusions can be reached:
The flexural behavior of UHPC beams can be obtained from a sectional analysis based on a linear strain distribution in a domain bounded by the UHPC compression strain limit, the UHPC localization strain limit, and the rupture strain of the reinforcing steel. The stresses on strained cross sections can be obtained from appropriate uniaxial stress-strain models for UHPC, which are different from those traditionally used for conventional concrete.
For flexural design, the UHPC uniaxial compression stress-strain behavior can be idealized by an elastic-perfectly plastic stress-strain model. The model requires three material properties: the modulus of elasticity, the compressive strength, and the strain at compression strength (compression strain limit). These properties can be obtained from compression tests performed on cylindrical specimens according to ASTM C1856 (ASTM 2017).
For flexural design, the tensile stress-strain response of UHPC can be idealized by an elastic-perfectly plastic or bilinear stress-strain relationship. The model requires three material property parameters: the cracking stress, the localization stress, and the localization strain. These properties can be obtained by executing a direct tension test on UHPC prismatic specimens.
The maximum moment capacity of UHPC beams can occur at crack localization when the strain in the extreme tensile layer is equal to the UHPC localization strain, or at compression crushing prior to localization when the strain in the extreme compression layer is equal to the UHPC ultimate compression strain. In each case, the opposing criteria must not be exceeded when the primary criteria is reached.
For members failing at crack localization, the nominal moment–sectional curvature diagram can be idealized by four key points corresponding to the initiation of flexural cracks, yielding of the bottom layer of reinforcement, crack localization, and concrete crushing. For members failing in compression prior to localization, the moment-curvature diagram is simplified into the two or three branches limited by concrete crushing, which can occur before or after the tensile yielding of the reinforcement.
After localization in flexure, beams made with UHPC are expected to develop one dominant crack. This behavior will cause the girder to hinge at the cross section of the localized crack, subjecting the tensile steel reinforcement crossing the localization crack to excessive strains and eventual rupture. For this reason, the postlocalization flexural capacity is not recommended for use in design. More research is needed to quantify the relationship between the rupture strain of the flexural reinforcement after localization, the opening of the localized crack, and the development length of the reinforcement in the vicinity of the localized crack.
The localization strain of UHPC attained during the pretensioned girder flexural test was approximately 24% greater than the localization strain obtained from direct tension testing, indicating that the primary tensile reinforcement may enhance the tensile strain capacity of the UHPC. More research is needed to refine the understanding of this behavior. Moreover, the compression strain limit attained during the test was approximately 20% larger than the ultimate compression strain specified in the simulations, suggesting that the UHPC in a structural element can undergo compression strains exceeding the strain at peak stress obtained from a uniaxial compression test.
Performing a capacity-based flexural design of UHPC girders that ignores the UHPC tensile resistance may incorrectly predict a concrete crushing failure mode at the top of the section, rather than localization of cracks at the bottom, and significantly overestimate the curvature ductility ratio, inadvertently resulting in a structure that does not demonstrate the desired ductility.
The results of the proposed design methodology were validated for a pretensioned girder tested by the authors and the results are in good agreement with the experimental data. The predicted-over-experimental flexural moment ratios were 0.97, 1.09, 0.99, and 0.92 at first crack, 80% of yield, yield, and crack localization, respectively. Additionally, the model nominal moment predictions for 40 prestressed and non-prestressed beams found in the literature are in good agreement with the experimental data.The average predicted-over-experimental ratio was 0.96 with a COV of 19.6% for all beams.

Notation

The following symbols are used in this paper:
d
distance from top of the section to the centroid of reinforcing steel;
Ec
modulus of elasticity of UHPC;
E¯c
average modulus of elasticity of UHPC;
Ep
modulus of elasticity of prestressing steel strands;
Es
modulus of elasticity of conventional steel;
fc
stress in extreme compression layer;
fc
compressive strength of UHPC;
f¯c
average compressive strength of UHPC;
fcrd,F
design cracking strength of UHPC as defined by Fig. 3.206 of NF P18-710;
fps
stress in prestressing strand layer;
fpu
ultimate tensile strength of prestressing strands;
f¯pu
average ultimate tensile strength of prestressing strands;
fpy
yield strength of prestressing strands;
f¯py
average yield strength of prestressing strands;
fs
stress in reinforcing steel layer;
fs,d
design stress of reinforcing steel as defined by the appropriate international design document;
fsy
yield stress of conventional steel;
ft,r
effective cracking stress of UHPC;
f¯t,cr
average effective cracking stress of UHPC;
ftd,C
design tensile strength of UHPC as defined by Fig. CA8.1.8a of CSA S6:19;
ftd,F
design tensile strength of UHPC as defined by Fig. 3.206 of NF P18-710;
ftd,S
design tensile strength of UHPC as defined by Section 2.4.2.3 of SIA 2052;
ft,lim
tensile strain limit under service loads;
ft,loc
localization stress of UHPC;
f¯t,loc
average localization stress of UHPC;
h
overall thickness or height of a member;
M
applied bending moment on UHPC girder;
Mc
flexural moment at crushing of UHPC at extreme compression layer;
Mcr
flexural moment at the first flexural crack;
Mexp
total experimental moment corresponding to applied moment, M, plus the self-weight moment, Mini;
Mini
bending moment induced by the self-weight of the UHPC girder and loading apparatus;
ML
flexural moment at crack localization, when the extreme tensile layer, εt, is equal to the localization strain of UHPC, εt,loc;
Mn
nominal flexural moment;
Mr
factored flexural moment;
Msl
baseline flexural moment for ductility considerations, calculated when the bottom layer of steel is equal to 80% the yield stress of steel reinforcement;
My
flexural moment at yielding of the extreme layer of reinforcement;
M1
maximum flexural moment as defined by Section CA8.1.8.4.3.1 of CSA S6:19 documents;
X
distance from extreme compression fiber to the neutral axis of the section;
α
reduction factor on the compressive strength reflecting the linearity limit of the material compressive stress-strain response;
γ
reduction factor applied on the tensile stress parameters (ft,cr and ft,loc) to account for variability in the material testing results;
ΔX
distance from extreme compression fiber to the neutral axis of the section of the intermediate strain profile in Fig. 13;
ΔεB
change in strain at midspan at extreme bottom fiber;
Δεcp1
change in strain in the concrete at midspan at the level of the first layer of strands;
Δεcp2
change in strain in the concrete at midspan at the level of the second layer of strands;
Δεcp3
change in strain in the concrete at midspan at the level of the third layer of strands;
Δεi
change in strain in UHPC layer i;
ΔεT
change in strain at midspan at extreme top fiber;
Δψ
change in sectional curvature taken as Δεt/ΔX;
εB
strain at midspan at extreme bottom fiber;
εB,ini
initial strain at midspan at extreme bottom fiber, corresponding to σB,ini;
εc
strain in extreme compression layer;
εcp
elastic compression strain limit taken as αfc/E;
εcp1
strain in the concrete at midspan at the level of the first (top) layer of strands;
εcp2
strain in the concrete at midspan at the level of the second layer of strands;
εcp3
strain in the concrete at midspan at the level of the third (bottom) layer of strands;
εcp1,ini
initial strain in the UHPC at the level of the first (top) layer of prestressing strands;
εcp2,ini
initial strain in the UHPC at the level of the second layer of prestressing strands;
εcp3,ini
initial strain in the UHPC at the level of the third (bottom) layer of prestressing strands;
εcu
strain at ultimate compressive strength of UHPC;
ε¯cu
average strain at ultimate compressive strength of UHPC;
εi
strain in UHPC layer i;
εi,ini
initial strain in UHPC layer i;
εpe
strain in the effective prestress;
εps
strain in the prestressing strand layer;
εpu
rupture tensile rupture strain of prestressing strands;
ε¯pu
average tensile rupture strain of prestressing strands;
εpy
yield strain of prestressing strands;
εp1
strain in the first (top) layer of steel strands;
εp2
strain in the second layer of steel strands;
εp3
strain in the third (bottom) layer of steel strands;
εp1,ini
initial strain in the first (first) layer of prestressing strands (at the star Mini);
εp2,ini
initial strain in the second layer of prestressing strands (strain at Mini);
εp3,ini
initial strain in the third (bottom) layer of prestressing strands (strain at Mini);
εs
strain in reinforcing steel layer;
εsu
rupture strain of conventional steel;
εsy
yield strain of conventional steel;
εT
strain at midspan at extreme top fiber;
εT,ini
initial strain at midspan at extreme top fiber, corresponding to σT,ini;
εt
strain in extreme tensile layer;
εt,cr
strain at effective cracking stress of UHPC;
εt,lim
design tensile strain limit of UHPC as defined by the appropriate international design document;
εt,loc
average localization strain of UHPC;
ε¯t,loc
average localization strain of UHPC;
μ
curvature ductility ratio defined as ψL/ψy;
μexp
experimental value of curvature ductility ratio;
μmin
specified minimum curvature ductility ratio;
μnt,sim
simulated value of curvature ductility ratio when the tensile resistance of UHPC is ignored;
μsim
simulated value of curvature ductility ratio;
ρ
reinforcement ratio of conventional steel;
σB,ini
initial stress at midspan at extreme bottom fiber;
σT,ini
initial stress at midspan at extreme top fiber;
ϕf
resistance factor for flexural moment;
ψ
sectional curvature defined as the ration of the compressive strain in the section over the depth of the neutral axis;
ψc
sectional curvature at crushing of UHPC at extreme compression layer;
ψcr
sectional curvature at the first flexural crack;
ψL
sectional curvature at crack localization, when the extreme tensile layer, εt, is equal to the localization strain of UHPC, εt,loc;
ψsl
baseline sectional curvature for ductility considerations, calculated when the bottom layer of steel is equal to 80% the yield stress of steel reinforcement;
ψy
sectional curvature at yielding of the extreme layer of reinforcement; and
ω¯c
average density of UHPC.

Appendix. Supplemental Background Information

Flexural Design of Concrete Members

Flexural design of concrete members in the United States is founded on mechanical models and sectional design methods specified in AASHTO LRFD BDS (AASHTO 2020) and ACI 318-19 (ACI 2019). In these documents, the analysis of the beam is based on the concepts of equilibrium and strain compatibility utilizing mechanical models to determine the stresses on strained cross sections. In compression, the stress-strain behavior of concrete at ultimate limit state may be simplified by a calibrated rectangular stress block. At ultimate loading, discrete longitudinal steel reinforcements are added and proportioned to carry all tensile stresses while ignoring the minimal tensile resistance offered by the concrete. When subjected to flexural loads, the tensile reinforcements span the concrete cracks and enable flexural tensile strains far greater than that at first concrete cracking. The ultimate capacity of the beam is commonly limited by the attainment of a flexural compression failure, defined when the strain in the concrete in compression reaches the assumed compression strain limit of 0.003.
Reinforced concrete flexural elements are described as (1) compression-controlled when the net tensile strain in the extreme tension steel at ultimate is less than a specified value, generally associated with the yield strain; (2) tension-controlled when the steel strain at failure is greater than a specified tensile strain limit value, e.g., 0.005 for non-prestressed reinforcement with a specified minimum yielding stress not exceeding 517 MPa (75 ksi) (AASHTO 2020); and (3) within the transition region when the net tensile strain in the steel at failure is between the specified thresholds of the compression- and tension-controlled sections. The steel tensile threshold value for the tension-controlled sections is expected to be sufficiently large to allow for excessive deformation and cracking, providing for a ductile behavior and ample warning of impending failure. In contrast, given the small tensile strains in compression-controlled sections, this failure mechanism may be brittle, with little warning of impending failure. Moreover, the failure of common compression members, i.e., columns, can lead to detrimental consequences to the stability of the entire structure. To address this issue, a reduced moment resistance factor is specified for compression-controlled sections, i.e., ϕf=0.75, compared to the resistance factor specified for tension-controlled sections, i.e., ϕf=0.90, for non-prestressed and prestressed components designed according to ACI 318-19 (ACI 2019), and ϕf=0.90 and 1.00 for non-prestressed components and prestressed components designed according to AASHTO LRFD BDS (AASHTO 2020), respectively.

International Recommendations for Flexural Design

In the French design standard, the resisting forces of the sections of a member subjected to flexure are calculated from a linear strain diagram included in a domain bounded by a compression strain limit, corresponding to UHPC crushing, and a reinforcing steel limit, corresponding to the rupture of reinforcing steel (AFNOR 2016a). For members with no steel reinforcing bars or prestressing strands, tensile strains are limited to the strain at peak tensile strength of UHPC. The cross-sectional forces are obtained by employing compression and tension mechanical stress-strain models to calculate the stresses corresponding to the assumed strains. The compression models include a detailed model based on parabolic representation of the stress-strain results obtained from compression tests and a simplified elastic-perfectly plastic model with reduced design stress and strain limits. The tension models include an elastic-perfectly plastic or bilinear stress-strain representation, limited by the strain at crack localization (hardening limit), for materials exhibiting a constant or increasing stress after cracking, respectively. The material compression parameters are obtained directly from uniaxial compression tests, while the tensile parameters are obtained indirectly from a prism flexure test using an inverse analysis technique. The inverse analysis back-calculates an approximation of the uniaxial tensile stress-strain response from the moment-deflection results of the prism test and includes specific procedures to infer the sectional curvatures of the beam corresponding to the measured deflections at each load step (AFNOR 2016b).
In the Swiss design recommendations, the ultimate bending resistance of a reinforced UHPC cross section is calculated by resolving the stresses corresponding to a linear strain diagram included in a domain bounded by the strains when the peak compression (crushing) and tensile (localization) strengths are reached (SIA 2016). The compression stress-strain model is linear elastic to failure, with the failure stress defined by the design compression strength obtained from compression testing of cylinders or cubes. In tension, a rectangular stress block is permitted with a stress equal to the design tensile strength (which is a proportion of the peak tensile strength) and extending to a distance of 0.9(hX) from the extreme tension fiber, where h is height of the beam section and X is the depth of the neutral axis. If the tensile model is used to describe a UHPC member with steel reinforcing bars or strands as the primary flexural tensile reinforcement, the tensile strain limit (strain at peak tensile strength) may be increased by a factor of 2.0 due to simultaneous action of the fibers and reinforcing steel; in this case, the design tensile strength is taken as a reduced value of the elastic limit (cracking stress). The uniaxial tensile stresses and strains of UHPC are obtained either from a direct tension test or from an inverse analysis of the load-deflection results of flexural prism tests. The inverse analysis method assumes a stress distribution in the section at peak force and uses simplified procedures to obtain the strain-hardening portion of the UHPC stress-strain response based on the measured vertical deflection and sectional curvatures of the prism during the test. The applicability of the Swiss design recommendation document is limited to UHPC materials exhibiting a uniaxial stress-strain response showing a continuous increase in stress after cracking until the peak stress is reached, i.e., strain-hardening behavior.
In the Canadian design recommendations, the design capacity of flexural elements composed of strain-hardening fiber-reinforced concrete (SH-FRC), such as UHPC, is calculated by determining the sectional forces corresponding to a linear strain diagram bounded by the SH-FRC crushing strain limit in compression and the rupture of steel reinforcement in tension (Annex A8.1 of CSA S6:19). The compression stress-strain model is linear elastic when the strain value in the extreme compression layer is less than the limiting strain at crushing. When the extreme compression fiber of the member equals the ultimate compression strain (i.e., crushing limit), the compression model can be taken as the traditional rectangular stress block, which is compatible with a triangular stress distribution for materials with compression strength values exceeding 120 MPa (17.4 ksi). The tensile model at the ultimate limit state is a rectangular distribution extending over cross-sectional layers where the tensile strains are less than or equal to the limiting tensile strain (i.e., strain at peak tensile resistance). The compression parameters are obtained from uniaxial compression tests, while the uniaxial tensile parameters (i.e., peak stress and strain) are obtained either from a direct tension test or from an inverse analysis of the load-deflection results of flexural prism tests. The inverse analysis technique is based on empirical approximations of the stresses at specific vertical deflections and crack openings recorded during the testing of a flexural prism of specified dimensions. The Canadian design document specifies a flexural resistance method based on sectional curvature ductility. The maximum flexural resistance M1 is taken as the maximum of the flexural moment (1) at yielding of the reinforcement in tension, (2) when the strain in the extreme tension fiber is equal to the limiting tensile strain, or (3) at concrete crushing. The curvature ductility ratio is defined as the ratio of the sectional curvature at M1 to the sectional curvature at the yielding of the tensile flexural reinforcement. If the ductility ratio is greater than 2.0, the flexural resistance of the member is taken equal to the moment at yielding of reinforcement. In cases where the curvature ductility is less than 2.0, the flexural resistance of the member must be taken as the maximum of 0.50×M1 or the flexural resistance calculated without considering the contribution of fibers. Fig. 1 presents a comparison of the cross-sectional stress conditions of UHPC members computed following the recommendations of the French, Swiss, and Canadian documents when the strain at the extreme tension layer is equal to the design tensile strain limit εt,lim and with elastic stresses in compression.

Data Availability Statement

Some or all data, models, or code that support the findings of this study are available from the corresponding author upon reasonable request. Some or all data, models, or code generated or used during the study are proprietary or confidential in nature and may only be provided with restrictions.

Acknowledgments

The research presented in this paper was funded by FHWA. The publication of this paper does not necessarily indicate approval or endorsement of the findings, opinions, conclusions, or recommendations either inferred or specifically expressed herein by FHWA or the United States Government. This research could not have been completed were it not for the dedicated support of the technical professionals associated with the FHWA Structural Concrete Research Program.

References

Aaleti, S., B. Petersen, and S. Sritharan. 2013. Design guide for precast UHPC Waffle Deck panel system, including connections. Washington, DC: Federal Highway Administration.
AASHTO. 2020. LRFD bridge design specifications. 9th ed. Washington, DC: AASHTO.
AASHTO. 2022. Standard method of test for uniaxial tensile response of ultra-high performance concrete. AASHTO T 397 Washington, DC: AASHTO.
ACI (American Concrete Institute). 2019. Building code requirements for structural concrete and commentary. ACI 318-19. Farmington Hills, MI: ACI.
AFNOR (Association Française de Normalisation). 2016a. National addition to the Eurocode 2—Design of concrete structures: Specific rules for ultra-high performance fibre-reinforced concrete (UHPFRC). Paris: AFNOR.
AFNOR (Association Française de Normalisation). 2016b. Ultra-high performance fibre-reinforced concrete—Specifications, performance, production, and conformity. Paris: AFNOR.
ASTM. 2014. Standard test method for static modulus of elasticity and poisson’s ratio of concrete in compression. ASTM C469/C469M. West Conshohocken, PA: ASTM.
ASTM. 2017. Standard practice for fabricating and testing specimens of ultra-high performance concrete. ASTM C1856/C1856M. West Conshohocken, PA: ASTM.
ASTM. 2018a. Standard specification for low-relaxation, seven-wire steel strand for prestressed concrete. ASTM A416/A416M. West Conshohocken, PA: ASTM.
ASTM. 2018b. Standard test method for compressive strength of cylindrical concrete specimens. ASTM C39/C39M. West Conshohocken, PA: ASTM.
ASTM. 2020. Standard test method and definitions for mechanical testing of steel products. ASTM A370. West Conshohocken, PA: ASTM.
Baby, F., P. Marchand, and F. Toutlemonde. 2014. “Shear behavior of ultrahigh performance fiber-reinforced concrete beams. I: Experimental investigation.” J. Struct. Eng. 140 (5): 04013111. https://doi.org/10.1061/(ASCE)ST.1943-541X.0000907.
Bencardino, F., L. Rizzuti, G. Spadea, and R. N. Swamy. 2010. “Experimental evaluation of fiber reinforced concrete fracture properties.” Composites, Part B 41 (1): 17–24. https://doi.org/10.1016/j.compositesb.2009.09.002.
Bolander, J. E., S. Choi, and S. R. Duddukuri. 2008. “Fracture of fiber-reinforced cement composites: Effects of fiber dispersion.” Int. J. Fract. 154 (1–2): 73–86. https://doi.org/10.1007/s10704-008-9269-4.
Chen, L., and B. A. Graybeal. 2012. “Modeling structural performance of ultrahigh performance concrete I-girders.” J. Bridge Eng. 17 (5): 754–764. https://doi.org/10.1061/(ASCE)BE.1943-5592.0000305.
Chen, S., R. Zhang, L. J. Jia, and J. Y. Wang. 2018. “Flexural behaviour of rebar-reinforced ultra-high-performance concrete beams.” Mag. Concr. Res. 70 (19): 997–1015. https://doi.org/10.1680/jmacr.17.00283.
CSA (Canadian Standard Association). 2019. Canadian highway bridge design code. S6:19. Torronto, ON: CSA.
Cunha, V. M. C. F., J. A. O. Barros, and J. M. Sena-Cruz. 2011. “An integrated approach for modelling the tensile behaviour of steel fibre reinforced self-compacting concrete.” Cem. Concr. Res. 41 (1): 64–76. https://doi.org/10.1016/j.cemconres.2010.09.007.
De Larrard, F., and T. Sedran. 2002. “Mixture-proportioning of high-performance concrete.” Cem. Concr. Res. 32 (11): 1699–1704. https://doi.org/10.1016/S0008-8846(02)00861-X.
Doyon-Barbant, J., and J. P. Charron. 2018. “Impact of fibre orientation on tensile, bending and shear behaviors of a steel fibre reinforced concrete.” Mater. Struct. 51 (6): 157. https://doi.org/10.1617/s11527-018-1282-0.
El-Helou, R., and B. Graybeal. 2019. “The ultra girder: A design concept for a 300-foot single span prestressed ultra high-performance concrete bridge girder.” In Proc., 2nd Int. Interactive Symp. on Ultra-High Performance Concrete. Ames, IA: Iowa State Univ.
El-Helou, R. G. 2016. “Multiscale computational framework for analysis and design of ultra-high performance concrete structural components and systems.” Ph.D. dissertation, Dept. of Civil and Environmental Engineering, Virginia Polytechnic Institute and State Univ.
El-Helou, R. G., Z. B. Haber, and B. A. Graybeal. 2022. “Mechanical behavior and design properties of ultra-high performance concrete.” ACI Mater. J. 119 (1). https://doi.org/10.14359/51734194.
El-Helou, R. G., I. Koutromanos, C. D. Moen, and M. Moharrami. 2020. “Triaxial constitutive law for ultra-high-performance concrete and other fiber-reinforced cementitious materials.” J. Eng. Mech. 146 (7): 04020062. https://doi.org/10.1061/(ASCE)EM.1943-7889.0001777.
Gowripalan, N., and I. R. Gilbert. 2000. Design guidelines for ductal prestressed concrete beams. Sydney: The University of New South Wales, VSL (Aust) Pty Ltd.
Graybeal, B. A. 2006a. Material property characterization of ultra-high performance concrete. McLean, VA: Federal Highway Administration.
Graybeal, B. A. 2006b. Structural behavior of ultra-high performance concrete prestressed I-girders. McLean, VA: Federal Highway Administration.
Graybeal, B. A. 2008. “Flexural behavior of an ultrahigh-performance concrete I-girder.” J. Bridge Eng. 13 (6): 602–610. https://doi.org/10.1061/(ASCE)1084-0702(2008)13:6(602).
Graybeal, B. A. 2009. Structural behavior of a prototype ultra-high performance concrete pi-girder. McLean, VA: Federal Highway Administration.
Graybeal, B. A., and F. Baby. 2013. “Development of direct tension test method for ultra-high- performance fiber-reinforced concrete.” ACI Mater. J. 110 (2): 177–186.
Graybeal, B. A., and B. Stone. 2012. Compression response of a rapid-strengthening ultra-high performance concrete formulation. McLean, VA: Federal Highway Administration.
Habel, K., M. Viviani, E. Denarié, and E. Brühwiler. 2006. “Development of the mechanical properties of an ultra-high performance fiber reinforced concrete (UHPFRC).” Cem. Concr. Res. 36 (7): 1362–1370. https://doi.org/10.1016/j.cemconres.2006.03.009.
Haber, Z., I. De La Varga, B. Graybeal, B. Nakashoji, and R. El-Helou. 2018. Properties and behavior of UHPC-class materials. McLean, VA: Federal Highway Administration.
Hasgul, U., K. Turker, T. Birol, and A. Yavas. 2018. “Flexural behavior of ultra-high-performance fiber reinforced concrete beams with low and high reinforcement ratios.” Struct. Concr. 19 (6): 1577–1590. https://doi.org/10.1002/suco.201700089.
Huang, H., X. Gao, L. Li, and H. Wang. 2018. “Improvement effect of steel fiber orientation control on mechanical performance of UHPC.” Constr. Build. Mater. 188 (Nov): 709–721. https://doi.org/10.1016/j.conbuildmat.2018.08.146.
Lepech, M. D., and V. C. Li. 2009. “Water permeability of engineered cementitious composites.” Cem. Concr. Compos. 31 (10): 744–753. https://doi.org/10.1016/j.cemconcomp.2009.07.002.
Li, C., Z. Feng, L. Ke, R. Pan, and J. Nie. 2019. “Experimental study on shear performance of cast-in-place ultra-high performance concrete structures.” Materials (Basel) 12 (19): 3254. https://doi.org/10.3390/ma12193254.
Magureanu, C., I. Sosa, C. Negrutiu, and B. Heghes. 2012. “Mechanical properties and durability of ultra-high-performance concrete.” ACI Mater. J. 109 (2): 177–184.
Maya Duque, L. F., I. De La Varga, and B. Graybeal. 2016. “Fiber reinforcement influence on the tensile response of UHPFRC.” In Proc., 1st Int. Interactive Symp. on UHPC. Ames, IA: Iowa State Univ.
Meade, T., and B. Graybeal. 2010. “Flexural response of lightly reinforced ultra-high performance concrete beams.” In Proc., 3rd fib Int. Congress. Washington, DC: Fédération International du béton.
NJDOT (New Jersey DOT). 2016. Design manual for bridges and structures. 6th ed. Trenton, NJ: NJDOT.
Oesterlee, C. 2010. “Structural response of reinforced UHPFRC and RC composite members.” Ph.D. dissertation, School of Architecture, Civil and Environmental Engineering, Swiss Federal Institute of Technology Lausanne.
PCI (Precast/Prestressed Concrete Institute). 2014. PCI bridge design manual. 3rd ed. Second Release. Chicago: PCI.
Peng, X., and C. Meyer. 2000. “A continuum damage mechanics model for concrete reinforced with randomly distributed short fibers.” Comput. Struct. 78 (4): 505–515. https://doi.org/10.1016/S0045-7949(00)00045-6.
Pros, A., P. Diez, and C. Molins. 2012. “Modeling steel fiber reinforced concrete: Numerical immersed boundary approach and a phenomenological mesomodel for concrete-fiber interaction.” Int. J. Numer. Methods Eng. 90 (1): 65–86. https://doi.org/10.1002/nme.3312.
Qiu, M., X. Shao, K. Wille, B. Yan, and J. Wu. 2020a. “Experimental investigation on flexural behavior of reinforced ultra high performance concrete low-profile T-beams.” Int. J. Concr. Struct. Mater. 14 (1): 5. https://doi.org/10.1186/s40069-019-0380-x.
Qiu, M., Y. Zhang, S. Qu, Y. Zhu, and X. Shao. 2020b. “Effect of reinforcement ratio, fiber orientation, and fiber chemical treatment on the direct tension behavior of rebar-reinforced UHPC.” Constr. Build. Mater. 256 (Sep): 119311. https://doi.org/10.1016/j.conbuildmat.2020.119311.
Radtke, F. K. F., A. Simone, and L. J. Sluys. 2010. “A computational model for failure analysis of fibre reinforced concrete with discrete treatment of fibres.” Eng. Fract. Mech. 77 (4): 597–620. https://doi.org/10.1016/j.engfracmech.2009.11.014.
Russel, H., and B. A. Graybeal. 2013. Ultra-high performance concrete: A state-of-the-art report for the bridge community. McLean, VA: Federal Highway Administration.
Schauffert, E. A., and G. Cusatis. 2012. “Lattice discrete particle model for fiber-reinforced concrete. I: Theory.” J. Eng. Mech. 138 (7): 826–833. https://doi.org/10.1061/(ASCE)EM.1943-7889.0000387.
Shao, Y., and S. L. Billington. 2019. “Utilizing full UHPC compressive strength in steel reinforced UHPC beams.” In Proc., 2nd Int. Interactive Symp. on Ultra-High Performance Concrete. Ames, IA: Iowa State Univ.
SIA (Swiss Society of Engineers and Architects). 2016. Recommendation: Ultra-high performance fibre-reinforced cement-based composites (UHPFRC) construction material, dimensioning, and application. Lausanne, Switzerland: SIA.
Skogman, B. C., M. K. Tadros, and R. Grasmick. 1988. “Flexural strength of prestressed concrete members.” PCI J. 33: 96–123.
Smith, J., G. Cusatis, D. Pelessone, E. Landis, J. O’Daniel, and J. Baylot. 2014. “Discrete modeling of ultra-high-performance concrete with application to projectile penetration.” Int. J. Impact Eng. 65 (Mar): 13–32. https://doi.org/10.1016/j.ijimpeng.2013.10.008.
Soranakom, C., and B. Mobasher. 2008. “Correlation of tensile and flexural responses of strain softening and strain hardening cement composites.” Cem. Concr. Compos. 30 (6): 465–477. https://doi.org/10.1016/j.cemconcomp.2008.01.007.
Soranakom, C., and B. Mobasher. 2009. “Flexural design of fiber-reinforced concrete.” ACI Mater. J. 106 (5): 461–469.
Stürwald, S. 2017. “Bending behaviour of UHPC reinforced with rebars and steel fibres.” In Proc., 2017 fib Symp.: High Tech Concrete: Where Technology and Engineering Meet, 473–481. Cham, Switzerland: Springer.
Walsh, K. K., N. J. Hicks, E. P. Steinberg, H. H. Hussein, and A. A. Semendary. 2018. “Fiber orientation in ultra-high-performance concrete shear keys of adjacent-box-beam bridges.” ACI Mater. J. 115 (2): 227–238. https://doi.org/10.14359/51701097.
Wille, K., and A. E. Naaman. 2013. “Effect of ultra-high-performance concrete on pullout behavior of high-strength brass-coated straight steel fibers.” ACI Mater. J. 110 (4): 451–462.
Yang, I. H., C. Joh, and B. S. Kim. 2010. “Structural behavior of ultra high performance concrete beams subjected to bending.” Eng. Struct. 32 (11): 3478–3487. https://doi.org/10.1016/j.engstruct.2010.07.017.
Yao, Y., X. Wang, K. Aswani, and B. Mobasher. 2017. “Analytical procedures for design of strain softening and hardening cement composites.” Int. J. Adv. Eng. Sci. Appl. Math. 9 (3): 181–194. https://doi.org/10.1007/s12572-017-0187-4.
Yin, H., K. Shirai, and W. Teo. 2019. “Finite element modelling to predict the flexural behaviour of ultra-high performance concrete members.” Eng. Struct. 183 (Mar): 741–755. https://doi.org/10.1016/j.engstruct.2019.01.046.
Yoo, D. Y., N. Banthia, and Y. S. Yoon. 2017. “Experimental and numerical study on flexural behavior of ultra-high-performance fiber-reinforced concrete beams with low reinforcement ratios.” Can. J. Civ. Eng. 44 (1): 18–28. https://doi.org/10.1139/cjce-2015-0384.
Yoo, D. Y., and Y. S. Yoon. 2015. “Structural performance of ultra-high-performance concrete beams with different steel fibers.” Eng. Struct. 102 (Nov): 409–423. https://doi.org/10.1016/j.engstruct.2015.08.029.

Information & Authors

Information

Published In

Go to Journal of Structural Engineering
Journal of Structural Engineering
Volume 148Issue 4April 2022

History

Received: Apr 13, 2021
Accepted: Sep 21, 2021
Published online: Jan 28, 2022
Published in print: Apr 1, 2022
Discussion open until: Jun 28, 2022

Authors

Affiliations

Research Structural Engineer, Genex Systems/Turner-Fairbank Highway Research Center, 6300 Georgetown Pike, McLean, VA 22101 (corresponding author). ORCID: https://orcid.org/0000-0003-0061-9439. Email: [email protected]
Benjamin A. Graybeal, Ph.D., M.ASCE https://orcid.org/0000-0002-3694-1369 [email protected]
P.E.
Team Leader, Bridge Engineering Research, FHWA Turner-Fairbank Highway Research Center, McLean, VA 22101. ORCID: https://orcid.org/0000-0002-3694-1369. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

  • Carbon Footprint between Steel-Reinforced Concrete and UHPC Beams, Journal of Structural Engineering, 10.1061/JSENDH.STENG-11449, 149, 3, (2023).
  • Shear Design of Strain-Hardening Fiber-Reinforced Concrete Beams, Journal of Structural Engineering, 10.1061/JSENDH.STENG-11065, 149, 2, (2023).
  • Monolithic and non-monolithic interface shear performance of ultra-high performance concrete, Engineering Structures, 10.1016/j.engstruct.2023.115667, 281, (115667), (2023).
  • Flexural strength of prestressed Ultra-High-Performance concrete beams, Engineering Structures, 10.1016/j.engstruct.2023.115612, 279, (115612), (2023).
  • Flexural behavior of small-sized I-shaped UHPC beams hybrid reinforced with steel plate and BFRP, Composite Structures, 10.1016/j.compstruct.2022.116595, 306, (116595), (2023).
  • A review of existing codes and standards on design factors for UHPC placement and fiber orientation, Construction and Building Materials, 10.1016/j.conbuildmat.2022.128308, 345, (128308), (2022).

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share