You are on page 1of 21

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/358242651

Flexural Behavior and Design of Ultrahigh-Performance Concrete Beams

Article in Journal of Structural Engineering · April 2022


DOI: 10.1061/(ASCE)ST.1943-541X.0003246

CITATIONS READS
13 708

2 authors:

Rafic El Helou Benjamin Allen Graybeal


National Research Council Associate at FHWA Turner-Fairbank Highway Research Ce… U.S. Department of Transportation
17 PUBLICATIONS 129 CITATIONS 153 PUBLICATIONS 5,626 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Dimensional Stability and Bond Performance of Grouted Connections View project

UHPC Material Characterization View project

All content following this page was uploaded by Benjamin Allen Graybeal on 31 January 2022.

The user has requested enhancement of the downloaded file.


Flexural Behavior and Design of
Ultrahigh-Performance Concrete Beams
Rafic G. El-Helou, Ph.D., M.ASCE 1; and Benjamin A. Graybeal, Ph.D., P.E., M.ASCE 2

Abstract: Beams made of ultrahigh-performance concrete (UHPC), a fiber-reinforced concrete with high compressive strength and tensile
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

strain-hardening characteristics, exhibit flexural behaviors that are different than those traditionally associated with steel-reinforced conven-
tional 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. DOI: 10.1061/(ASCE)ST.1943-541X.0003246. This work is made available under the terms of
the Creative Commons Attribution 4.0 International license, https://creativecommons.org/licenses/by/4.0/.
Author keywords: Ultrahigh-performance concrete; Ultrahigh-performance concrete (UHPC) beam; Flexural design model; Prestressed
girders; Strain-hardening fiber-reinforced concrete.

Introduction material volumes, load demands, girder depth, labor and construc-
tion time, equipment needs, and shipping and handling needs,
Ultrahigh-performance concrete (UHPC) is an emerging class of along with the potential elimination of girder lines or intermediate
cementitious composites, commonly proportioned using particle piers. However, the use of UHPC in primary structural components
packing theory and reinforced with discontinuous steel fibers, that in the United States has been generally limited due to its increased
offers superior mechanical and durability characteristics compared initial cost and the lack of formal design guidelines that allow for
to conventional and other fiber-reinforced concretes (De Larrard incorporating UHPC’s tensile ductility in design.
and Sedran 2002; Graybeal 2006a; Habel et al. 2006; Lepech and Flexural design of concrete members in the United States is
Li 2009; Bencardino et al. 2010; Magureanu et al. 2012; Wille and founded on mechanical models and sectional design methods speci-
Naaman 2013; El-Helou 2016). A typical UHPC-class material fied in the AASHTO Load and Resistance Factor (LRFD) Bridge
exhibits a compressive strength exceeding 124 MPa (18.0 ksi) Design Specifications, henceforth referred to as AASHTO LRFD
and a ductile strain-hardening tensile behavior with a cracking BDS (AASHTO 2020), and the American Concrete Institute (ACI)
strength above 5.0 MPa (0.73 ksi), sustained to a postcracking ten- 318-19 Building Code Requirements for Structural Concrete, hence-
sile strain greater than 0.0025 (Russel and Graybeal 2013; Haber forth referred to as ACI 318-19 (ACI 2019). In these documents, the
et al. 2018; El-Helou et al. 2022). When used in beams, UHPC can minimal tensile resistance of concrete is ignored, the discrete longi-
offer substantial tensile resistance, reducing the beam’s cross- tudinal steel reinforcements must be added and proportioned to carry
sectional dimensions, the weight of the superstructure, and the need all tensile stresses, and flexural failure is defined at concrete crushing
for traditional steel reinforcement or prestressing strands as com- when the compression strain in the concrete reaches 0.003. More-
pared to conventional concrete (Graybeal 2006b; El-Helou and over, the guidance for flexural effects in AASHTO LRFD BDS
Graybeal 2019). These features can lower costs by reducing (AASHTO 2020) is limited to concretes having a design compres-
sive strength less than or equal to 103 MPa (15 ksi). More informa-
1
Research Structural Engineer, Genex Systems/Turner-Fairbank High- tion on the flexural design of concrete members according to
way Research Center, 6300 Georgetown Pike, McLean, VA 22101 (corre- AASHTO LRFD BDS (AASHTO 2020) and ACI 318-19 (ACI
sponding author). ORCID: https://orcid.org/0000-0003-0061-9439. Email: 2019) can be found in the Appendix.
rafic.elhelou.ctr@dot.gov The mechanical behavior of UHPC is significantly different
2
Team Leader, Bridge Engineering Research, FHWA Turner-Fairbank from conventional concrete, necessitating substantive changes in
Highway Research Center, McLean, VA 22101. ORCID: https://orcid.org the common understanding of the flexural behavior. Due to the in-
/0000-0002-3694-1369. Email: benjamin.graybeal@dot.gov
creased UHPC compressive strength, the sustained UHPC post-
Note. This manuscript was submitted on April 13, 2021; approved on
September 21, 2021; published online on January 28, 2022. Discussion cracking tensile resistance, and the reduced development length
period open until June 28, 2022; separate discussions must be submitted of discrete reinforcements embedded in UHPC, beams made with
for individual papers. This paper is part of the Journal of Structural En- UHPC and reinforced with steel bars or prestressing strands often
gineering, © ASCE, ISSN 0733-9445. fail after the formation of a single localized crack, initiated by

© ASCE 04022013-1 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


the pullout of the crack-bridging fibers and followed by the subsequent test (i.e., εt;lim ¼ 2εt;loc ). Fig. 1 presents a comparison of the
rupture of the tensile reinforcement (Graybeal 2008; Meade and cross-sectional stress conditions of UHPC members computed fol-
Graybeal 2010; Yang et al. 2010; Yoo and Yoon 2015; Stürwald lowing the recommendations of the French, Swiss, and Canadian
2017; Yoo et al. 2017; Chen et al. 2018; Hasgul et al. 2018; Shao documents when the strain at the extreme tension layer is equal to
and Billington 2019). In this mode of failure, the peak flexural mo- the UHPC tensile strain limit εt;lim and with elastic stresses in com-
ment is attained at the initiation of the localized crack, at which point pression. Note that εt;lim is equal to εt;loc in the French and Canadian
the strain in the extreme tension fiber is equal to the crack localization models and 2εt;loc in the Swiss model. In Fig. 1, ftcr;F and ftd;F are
strain of UHPC (tensile strain limit) and the compressive strains are the design cracking and tensile strength as specified by the French
well below the crushing strain limit. After localization, the loss of the document (Fig. 3.206 of NF P18-710). Similarly, ftd;S and f td;C
UHPC fiber-bridging resistance mechanism leads to a decrease in are the design tensile stresses computed according to the Swiss
the flexural resistance of the member at the localized section, and (Section 2.4.2.3 of SIA 2052) and Canadian (Fig. CA8.1.8a of
in many cases, the load and deformation sustained by the member CSA S6:19) documents, respectively. The French and Canadian
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

prior to localization will generate a significant increase in curvature standards allow for tensile strains beyond the UHPC tensile strain
at the localized section, leading to the local straining of the tensile limit εt;lim , shown in Fig. 1, in which case the bending moment can
reinforcement and the hinging of the beam around the localized crack. be computed when the extreme compression strain is equal to the
These observations accentuate the need for a novel design and analy- UHPC crushing limit. In this instance, the contribution of UHPC
sis framework that accounts for the enhanced mechanical properties beyond the tensile strain limit is neglected and the compression
of UHPC and captures the new flexural failure modes expected with stress-strain model is represented by an elastic-perfectly plastic
this class of materials. Using existing conventional concrete specifi- model according to the French standard or a rectangular stress
cations to design UHPC beams, in which the concrete tensile resis- block according to the Canadian recommendation. More informa-
tance is ignored in beam capacity calculations and the flexural tion on the UHPC mechanical models and flexural design method-
strength is governed by reinforcement yielding and concrete compres- ologies in each of the French, Swiss, and Canadian documents
sion failure, can result in formulaic simplifications and assumptions can be found in the Appendix. Although these models can predict
that are not consistent with the behaviors of a UHPC member. aspects of the flexural behavior of a UHPC beam, they do not
Within the international community, a few recommendation or highlight the differences in the behavior before and after crack
specification documents for the design of UHPC structural mem- localization. Because these models allow postlocalization tensile
bers have been published. Of note are the French, Swiss, and Cana- straining of the UHPC, excessive local straining of the tensile
dian UHPC documents [NF P18-710 (AFNOR 2016a); SIA 2052 reinforcement and hinging of the beam at the localized crack
(SIA 2016); CSA S6:19 (CSA 2019)]. In each of these documents, can occur in some instances; these behaviors are not recognized
the flexural methodology is based on sectional analysis utilizing the in the design methodologies.
concepts of equilibrium and strain compatibility while using new A few additional flexural design models for UHPC members
UHPC mechanical models to determine the stresses on strained can be found in the literature. A number of these models focus on
cross sections. These new mechanical models for UHPC replace either developing phenomenological constitutive models utilizing
or modify existing models for flexural design of conventional con- smeared crack finite-element model (FEM) approaches (e.g., Peng
crete members in their respective jurisdictions and account for the and Meyer 2000; Radtke et al. 2010; Cunha et al. 2011; Pros et al.
UHPC postcracking tensile resistance afforded by the discontinu- 2012; Li et al. 2019; El-Helou et al. 2020), applying existing dam-
ous fiber reinforcement. age plasticity models within commercially available FEM software
In the French and Canadian design standards, the flexural mo- (e.g., Chen and Graybeal 2012; Yin et al. 2019), or formulating
ment capacity of a steel-reinforced UHPC beam is calculated from lattice or lattice discrete particle model approaches (Bolander et al.
the sectional forces at equilibrium corresponding to a linear strain 2008; Schauffert and Cusatis 2012; Smith et al. 2014; El-Helou
diagram included in a domain bounded by a compression strain 2016). Although some of these models were validated to capture
limit, corresponding to UHPC crushing, and a reinforcing steel the behavior of UHPC flexural components accounting for the ef-
limit, corresponding to the rupture of reinforcing steel. In contrast, fect of fiber orientation and tensile ductility, they are not suitable for
the ultimate flexural moment in the Swiss design recommendation direct implementation into design guidelines or common design
is calculated when the tensile strain in the section reaches the calculation processes.
UHPC tensile strain limit εt;lim , taken equal to double the value Realizing the need for simplified analytical procedures, some
of the crack localization strain εt;loc , as obtained from a material researchers focused on sectional analysis approaches, applying

Beam Strains Stresses Stresses Stresses


c fc fc fc

compression X
M
neutral
d axis
h ftcr,F
tension t,cr

0.9(h-X)
s fs,d fs,d fs,d
t,lim ftd,F ftd,S ftd,C
(a) (b) (c)

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.

© ASCE 04022013-2 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


the concepts of strain compatibility and force equilibrium and pro- 813 mm (32 in.)
posing new uniaxial stress-strain relationships for UHPC. For in- 117.5 mm (4 85 in.)
267 mm 50.8 mm (2 in.)
stance, authors have suggested that the compression stress-strain 22.2 mm
(10 21 in.)
curve might be idealized with a linear relationship (e.g., Graybeal 50.8 mm (2 in.)
( 87 in.)
2008; Aaleti et al. 2013), elastic-perfectly plastic relationship 889 mm 178 mm 406 mm
(35 in.) (7 in.) (16 in.)
(e.g., Soranakom and Mobasher 2009; Yao et al. 2017), or trilinear
88.9 mm (3 21 in.)
relationship including an elastic-perfectly plastic and descending 229 mm
88.9 mm (3 21 in.)
branch to failure (e.g., Gowripalan and Gilbert 2000). In tension, (9 in.) 76.2 mm (3 in.)
the stress-strain curve has been idealized with an elastic-perfectly 127 mm (5 in.)
plastic relationship (e.g., Graybeal 2008; Aaleti et al. 2013), a tri- 813 mm (32 in.)
linear relationship with an elastic- perfectly plastic and descending indicates a 17.8 mm (0.7 in.) diameter steel strand, prestressed to a force of
265 kN (59.5 kips); strands are placed at 50.8 mm (2 in.) center-to-center and in
branch to failure (e.g., Gowripalan and Gilbert 2000), or a trilinear three layers located at 50.8 mm (2 in.), 101.6 mm (4 in.), and 838 mm (33 in.)
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

relationship with biaxial and plastic branches (e.g., Soranakom and from bottom of section.
Mobasher 2008). In many of the existing studies, the uniaxial mod-
els, particularly the tension models, were obtained from inverse Fig. 2. Cross-sectional dimensions and reinforcement details of the
analysis of material testing of small-scale beams subjected to flex- tested girder.
ure, and the recommendations were limited to the tested specimens
and the behaviors of the materials investigated.
The presented considerations indicate the need for unified
girder’s cross-sectional shape was based on the Precast Concrete
mechanical models that are inferred from independent uniaxial
Economical Fabrication (PCEF) Committee recommended sec-
material testing methods and utilized in a flexural design method-
tions, with geometrical modifications to reduce the thickness of the
ology that is validated with large-scale experimental tests. Such a
bottom flange and the width of the top flange. The standard draw-
process is essential to determine whether material-level perfor-
ings for the original PCEF sections can be found in the New Jersey
mance metrics, which are typically available during the design
DOT Design Manual for Bridges and Structures (NJDOT 2016).
phase, can be relied upon to accurately determine the flexural
The cross-sectional dimensions and reinforcement details of the
capacity and failure mechanisms of steel-reinforced beams made
tested girder are shown in Fig. 2. The girder had a top and bottom
with UHPC-class materials. This manuscript addresses this need
flange width of 813 mm (32 in.), a web width of 178 mm (7 in.),
through the displacement-controlled testing of a full-scale pre-
and a height of 889 mm (35 in.); it was 18.90 m (62 ft) long and
stressed UHPC bridge girder to failure and the validation of a
contained a total of 26 17.8-mm- (0.7-in.)-diameter straight steel
mechanics-based flexural design methodology that can capture
strands, with 24 strands located in the bottom bulb and two strands
the failure modes of steel-reinforced UHPC beams while relying
in the top bulb, as indicated in Fig. 2. No mild steel reinforcement
on material parameters that are available during the design phase.
bars in the longitudinal or transverse directions were included in the
The experimental work focused on capturing specific flexural girder. The 17.8-mm- (0.7-in.)-diameter strands were Grade 1860-
behaviors at first cracking, yielding of prestressing strands, moment MPa (270-ksi) low-relaxation, seven-wire strands [ASTM A416/
capacity at the development of a localized crack, and rupture of the A416M (ASTM 2018a)] with a cross-sectional area of 189.7 mm2
strands. The flexural design framework is founded on the concepts (0.294 in:2 ) and a minimum ultimate strength of 1,860 MPa (270
of equilibrium and strain compatibility and is proposed for beams ksi); they were pretensioned to 75% of their ultimate strength and
made with UHPC and reinforced with conventional steel reinforc- placed at a minimum spacing of 50.8 mm (2 in.) center-to-center.
ing bars and/or prestressing strands. The method utilizes existing The first, second, and third layers of strands were located at
UHPC mechanical behavior models obtained from uniaxial 50.8 mm (2 in.), 101.6 mm (4 in.), and 838 mm (33 in.) from the
material tests, accounts for the flexural behaviors and expected fail- bottom of the section, respectively. The gross cross-sectional prop-
ure modes associated with UHPC members, and limits the tensile erties of the girder are an area of 3.72 × 105 mm2 (577 in:2 ), a
strain in beam sections to the UHPC strain at crack localization to moment of inertia about the centroid of 3.92 × 1010 mm4 (9.42 ×
avoid hinging of the beam around the localized crack. The ductility 104 in:4 ), and a centroid located at 432 mm (17 in.) from the bottom
of UHPC beams is also addressed through a sectional-curvature of the section.
approach that derives a resistance factor based on a threshold cur-
vature ductility ratio. The flexural design methodology is tailored to
generally align with existing bridge design specifications in the Mix Design and Girder Fabrication
United States (AASHTO 2020) in support of ongoing efforts to The UHPC product used in the girder was proprietary and was sup-
develop UHPC structural design guidance. The capability of the plied in three primary constituents, namely, the preblended powder
proposed method to predict the flexural behavior is verified by containing the granular constituents (e.g., cement, silica fume,
comparing the experimental data, obtained from the flexural test ground quartz, and fine sand), liquid admixtures, and fibers. The
described in this paper and other tests from the literature, to the mix design consisted of 2,182 kg=m3 (3,678 lb=yd3 ) of preblended
analytical predictions, relying on input material parameters ob- powder, 166 kg=m3 (280 lb=yd3 ) of water, 73.4 kg=m3 (124 lb=yd3 )
tained from independent material tests whenever available. 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-
Experimental Program 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
Test Specimen
constituents were blended first, then the water and superplasticizers
To conduct the experimental program described in this paper, one were added gradually. After the mixture turned from a powdered
pretensioned bulb-tee bridge girder made with UHPC was tested state into a viscous fluid, the fibers were dispersed into the mix.
under displacement-controlled loading in flexure to failure. The Before casting of each specimen, a flow table test was performed

© ASCE 04022013-3 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


per ASTM C1856 (ASTM 2017) to ensure a flowable consistency compression tests on two cylindrical specimens (one specimen per
corresponding to a spread equal to or greater than 203.2 mm (8 in.). batch) having a diameter of 76.2 mm (3 in.) and height of 152.4 mm
Two batches of UHPC of approximately 3.44 m3 (4.50 yd3 ) (6 in.). The cylinders were tested by the precast plant quality control
were mixed to obtain the total volume of fresh UHPC needed to crew according to ASTM C39 (ASTM 2018b) with modifications
make the girder. The fresh UHPC of the first batch, Batch A, filled listed in ASTM C1856 (ASTM 2017).
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. Test Setup and Instrumentation
The placement of fresh UHPC was performed from one stationary The 18.90-m- (62-ft)-long girder was tested in flexure with a span
discharge point located in the middle of the girder form. The of 18.29 m (60 ft) at the Federal Highway Administration (FHWA)
material flowed to each end of the form as it filled the bottom bulb; Turner-Fairbank Highway Research Center (TFHRC) structural
it then rose upward through the web height before starting to flow testing facility. The age of the girder at the time of the test was
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

again laterally as it filled the top bulb. The discharge point was then 155 days after casting. An overview of the test setup is shown in
moved sequentially along the girder length to top off the filling of Fig. 3. The girder was supported by a roller support at one end
the form. Minor external vibration was applied to the sides of the (west side) and by a partial pin support assembly bearing on a load
form to facilitate the release of entrapped air. The girder was then cell and a hydraulic jack at the other end (east side); it was loaded in
covered with a plastic sheet and cured at outdoor ambient temper- a four-point bending configuration with the middle reaction points
atures. The placement occurred during the month of June 2019 in each located at 0.46 m (1.5 ft) from the midspan, as shown in
which the average daily temperature ranged between 22°C (72°F) Fig. 4. The roller and partial pin supports each consisted of a
and 32°C (89°F). Five days after casting, the forms were removed 152-mm-(6-in.)-diameter steel cylinder inserted between two steel
and the strands were detensioned. The compressive strength bearing plates having a thickness of 38.1 mm (1.5 in.), a width of
required for detensioning was 96.5 MPa (14 ksi); the fabricator 0.30 m (1 ft), and a length equal to the width of the bottom flange.
elected to wait until the average compressive strength was 130.4 MPa The bearing plates of the pin support were grooved to prevent
(18.9 ksi). The compressive strength was determined by performing rolling, while the plates at the roller support were flat. To relieve

greased
reaction frame spreader beam spherical
bearing
spreader beam
spherical
west loadcells spherical bearings
bearing
west load cells
transfer plate east load
cell

top LVDTs

loading bottom LVDTs


jack
roller support

(a) (b)

Fig. 3. Overview of the experimental test setup showing (a) bridge girder and reaction frame; and (b) instrumentation at midspan.

0.46 m
(1.5 ft) centerline of reaction frame
east reaction
8.69 m west reaction 0.91 m spherical bearing
(28.5 ft) west load cells (3 ft) east load cell
transfer plate

top LVDTs: T1 T2 T3 T4 T5

west end east end

bearing plate bottom LVDTs: B1 B2 B3 B4 B5 bearing plate and PTFE sheet


roller support reaction pin support
7.01 m 1.83 m 0.61 m 1.37 m 1.22 m P
load cell
(23.0 ft) (6.0 ft) (2.0 ft) (4.5 ft) (4.0 ft)
loading jack
wire potentiometers: WP1 WP2 WP3 WP4 WP5 WP6
18.29 m
0.30 mm (1.0 ft) (60.0 ft) 0.30 m (1.0 ft)

Fig. 4. Test setup and instrumentation plan.

© ASCE 04022013-4 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


frictional forces on the pin support generated by the loading jack girder vertical displacement (WP6) of 216 mm (8.5 in.). The second
line of action relative to the inclination of the bearing location, an stage was halted when the jack load was 812 kN (182.1 kips) with a
additional steel bearing plate resting on a polytetrafluoroethylene girder vertical displacement (WP6) of 461 mm (18.1 in.). The third
(PTFE) sheet was included within the partial pin support assembly, and last stage subjected the girder to a maximum jack load of
as shown in Fig. 4. When the hydraulic jack applied load and de- 854 kN (192 kips), which occurred at a vertical displacement
formation onto the east end of the girder, two reaction points near (WP6) of 517 mm (20.4 in.), on its way to complete failure (strand
the midspan resisted the loading. These reaction points consisted of rupture) at a jack load of 761 kN (171 kips) and girder vertical
a pair of steel transfer plates reacting against the reaction frame displacement (WP6) of 610 mm (24.0 in.). Note that the stroke
through a set of three load cells with spherical bearing plates (two of the loading jack was reset once during the second test stage
on the west side and one on the east side of the midspan), a spreader and twice during the third test stage. The reset was performed using
beam, and a greased spherical bearing, as shown in Fig. 3(b). Each additional hydraulic jacks to hold the applied load while the stroke
of the middle transfer plates had a thickness of 102 mm (4 in.) and a of the loading jack was being retracted and spacer plates inserted
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

width of 0.30 m (1 ft). All transfer and bearing plates were grouted between the jack and the girder.
to the girder’s top or bottom flanges and were long enough to fully
support the width of the girder. Material Characterization
In addition to the four load cells, the girder was instrumented
with six wire potentiometers (WPs) and 13 linear variable displace- A key objective of this study was to evaluate the predicted flexural
ment transducers (LVDTs). The WPs (WP1–WP6) measured the response of the tested girder utilizing a proposed design method-
vertical deflection at selected locations along the span, as shown ology in which the mechanical properties had been obtained from
in Fig. 4. The strain profile throughout the test was captured by independent material tests. For this reason, companion specimens
mounting a pair of LVDTs, one at the top and one at the bottom were cast from each of the two batches used to make the girder to
assess the UHPC’s density, compression, and tension properties at
of the girder, in five regions (Regions 1–5) covering the middle
the time of the test. The material specimens were demolded at the
2.44 m (8 ft) of the span where monitoring of flexural cracking
time of detensioning of strands and stored in an ambient environ-
was desired. Figs. 3(b) and 4 show the top (T1–T5) and bottom
ment, generally alongside the girder, until the time of the test.
(B1–B5) LVDTs mounted on the girder in each of the five regions,
The average results of the material property tests are summarized
with Region 3 being the constant moment region. These LVDTs
in Table 1. The average density ω̄c, modulus of elasticity Ēc, com-
measured the average longitudinal displacement that can be trans-
pressive strength f̄c0 , and strain at compressive strength ε̄cu of each
formed into average strain by dividing the measured data by the
UHPC batch were obtained from three cylindrical specimens tested
LVDT gauge length. The top LVDTs had a gauge length of according to ASTM C1856 (ASTM 2017). The cylindrical speci-
610 mm (24 in.) and were mounted on one side of the top flange mens had a diameter of 76.2 mm (3 in.) and a height of 152.4 mm
at a vertical distance below the top extreme fiber of 25.4 mm (1 in.) (6 in.). The average tensile parameters, namely, the effective cracking
for LVDTs T1, T3, and T5, and of 76.2 mm (3 in.) for LVDTs T2 stress f̄t;cr , the localization stress f̄ t;loc , and the localization strain
and T4. The bottom LVDTs (B1–B5) had a gauge length of ε̄t;loc , were obtained from prismatic specimens tested in uniaxial ten-
584 mm (23 in.) and were mounted on the bottom surface of sion according to AASHTO T 397 (AASHTO 2022), which is based
the girder at a vertical distance of 50.8 mm (2 in.) below the girder. on the test method by Graybeal and Baby (2013). The tensile test
Finally, the slippage of the strands during the test was monitored captures the tensile load and associated strain over a 101.6-mm
through a set of three LVDTs mounted on three bottom-layer (4-in.) gauge length during a fixed-end, uniaxial displacement-
strands on the west end of the girder. controlled test. The prismatic specimens had a length of 432 mm
The load was applied by a hydraulic jack in 44.5-kN (10-kip) (17 in.) and a square 50.8 mm × 50.8 mm (2 in: × 2 in:) cross sec-
increments until the girder began to sustain inelastic damage and tion. The tensile parameters reported in Table 1 were obtained from
exhibit a reduced flexural stiffness. The loading was then switched the individual stress-strain curves of three and four prisms cast from
to displacement control at a jack displacement rate of between Batches A and B, respectively. In determining the individual results,
1.27 mm=min (0.05 in:=min) and 7.62 mm=min (0.3 in:=min). the effective cracking stress ft;cr was taken as the stress at the in-
Periodically throughout the test, unloading/reloading cycles of tercept of a line with a slope equal to the elastic modulus and a strain
89 kN (20 kip) were performed to measure the residual stiffness offset of 0.02% (Haber et al. 2018; El-Helou et al. 2022). The locali-
of the girder. The loading continued until girder failure, which zation stress f t;loc and strain, εt;loc , were visually determined as the
was defined as the formation of a dominant crack triggered by the first point in the stress-strain plot where the stress decreases contin-
pullout of the fiber reinforcement and the subsequent rupturing of uously with increasing strain. This point indicates the onset of fiber
the prestressing strands. Given that the girder deformation ex- pullout and the accumulation of the deformation into a single dom-
ceeded the stroke of the loading jack, the test was conducted by inant crack. The average uniaxial stress-strain curves for each batch
loading the girder in three separate test stages. At the end of each are shown in Fig. 5(a).
stage, the girder was completely unloaded to allow for the neces- The stress-strain relationships for the prestressing strands
sary preparations to reset the loading apparatus. The first test stage were experimentally determined from six strand pullout tests
was halted when the load at the jack was 547 kN (123 kips) with a performed according to ASTM A370 (ASTM 2020) and are plotted

Table 1. Average density and uniaxial compressive and tensile parameters for each UHPC batch at the time of the flexural test
Density Compression parameters Tension parameters
Batch
3 3
ID ω̄c [kg=m (lb=ft )] Ēc [GPa (ksi)] f̄c0 [MPa (ksi)] ε̄cu f̄t;cr [MPa (ksi)] f̄ t;loc [MPa (ksi)] ε̄t;loc
A 2,336 [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%]
B 2,323 [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.

© ASCE 04022013-5 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


14 2.0 2,000
275
1,800 Rupture
Batch B 1.8
12 Batch A points 250
1.6 1,600 Strands uniaxial
225
model
10

Tensile Stress (MPa)

Tensile Stress (MPa)


1,400

Tensile Stress (ksi)


1.4 200

Tensile Stress (ksi)


1.2 1,200 Experimental data 175
8 UHPC tensile
(6 strand tests)
model 1,000 150
1.0
6 125
0.8 800
100
4 0.6 600
Girder Cross Section 75
Batch B 0.4 400
50
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

2 Individual tests
Batch A 0.2 200
Average 25
0 0 0 0
0 0.002 0.004 0.006 0.008 0.01 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
(a) Uniaxial Strain (b) Uniaxial Strain

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.

in Fig. 5(b). The strain data were continuously collected until a portions at the transitions between the three test stages removed for
uniaxial strain of at least 0.03. The yield strength was measured clarity. The presented applied moment M is equal to the east and
at 1.0% extension under load and the total elongation value was west load cells readings multiplied by the flexural span of 8.69 m
determined by adding the 1.0% yield extension to the percent ex- (28.5 ft) and does not include the initial moment at the midspan
tension between the jaws gripping the strand at rupture, as specified M ini induced by the self-weight of the girder and the loading ap-
by ASTM A416 (ASTM 2018a). The strands had an average yield paratus (e.g., spreader beam, transfer plates, load cells). The initial
strength f̄ py of 1,690 MPa (246 ksi), an average ultimate tensile moment was equal to 467 kN-m (345 kip-ft) and was calculated
strength f̄pu of 1,932 MPa (281 ksi), and an average rupture strain based on the difference in the load cell readings before and after
ε̄pu of 0.0674. The coefficient of variation (COV) was 0.9% for f̄ py the girder and loading apparatus were installed. The east and west
and f̄ pu , and 5.1% for ε̄pu. The strands modulus of elasticity Ep reaction point deflections were calculated as the vertical displace-
was assumed to be 196.5 GPa (28,500 ksi). Based on these results, ment near the east and west load cells (WP2 and WP3) with respect
the strands conform to the mechanical property thresholds of to a reference line connecting the east and west supports. The data
ASTM A416 for Grade 1860 MPa (270 ksi) strands. 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 ap-
plied moment versus the average deflection behavior of the con-
Test Results, Analysis, and Observations stant moment region and was obtained by averaging the load
The applied moment versus deflection curves at the east and west and displacement results at the east and west reaction points
point loads are shown in Fig. 6(a), with the unloading and reloading throughout the test.

Deflection (in.)
0 1 2 3 4 5 6 7 8 9 10 11
8,000
Tension Compression 5,500
Applied Moment at Midspan (kN-m)

Applied Moment at Midspan (kip-ft)

7,000 5,000
East reaction
6,000 West reaction * * * 4,500
Average Test stage 3 Deformation localized into a single crack
(strain values reflect extension within 4,000
5,000 the gauge length of region 3) 3,500
Bottom of section Top
4,000 Test stage 2 3,000
of section
2,500
3,000
Linearity limit (first crack)
2,000
2,000 80% of yield stress in bottom layer of strands Region 1
1,500
Yielding of bottom layer of strands Region 2
Onset of strain localization (ultimate moment) Region 3 1,000
1,000 * Localized crack Region 4
Rupture of bottom strands 500
Test stage 1 Region 5
0 0
0 50 100 150 200 250 300 -0.03 -0.025 -0.02 -0.015 -0.01 -0.005 0 0.005
Deflection (mm) Change in Longitudinal Strain
(a) (b)

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).

© ASCE 04022013-6 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


The average midspan moment-deflection response of the girder The pair of top and bottom LVDTs that were installed in each of
was linearly elastic until softening started to occur at an applied the five instrumented regions of the girders (Fig. 4) were used to
moment of approximately 3,655 kN-m (2,695 kip-ft) and a deflec- create the average linear strain profiles in these regions throughout
tion of 65.9 mm (2.60 in.), indicating the initiation of flexural the duration of the test (assuming plane sections before bending
cracking in the girder. As the load continued to increase, the girder remain plane after bending). Each strain profile was then utilized
continued to sustain inelastic damage and exhibited a gradual to determine the change in concrete strains at extreme top (com-
decrease in the flexural stiffness. This behavior was likely due pression) and bottom (tension) fibers within the respective region
to the development of tight and closely spaced cracks within the as a function of the average applied moment at the midspan, as
tensile zone at the bottom half of the section. These cracks were shown in Fig. 6(b). The relationships in Fig. 6(b) confirm that, after
periodically investigated by halting the loading protocol during the localized crack initiated within the constant moment region
the first two test stages, up to an applied moment of 6,920 kN-m (Region 3) at the peak moment, the deformation of the cross section
(5,104 kip-ft), and by spraying the girder’s bottom flange and at the localized crack increased, as evidenced by the strain measure-
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

web within the constant moment region with an evaporative liquid ments at the top and bottom of the section, while the neighboring
(denatured alcohol). The liquid briefly penetrates the cracks while cracks closed as the applied load decreased. Note that the longitu-
evaporating from the girder’s surface, temporarily making the dinal deformation after the formation of a localized crack should be
cracks visible. However, none of the flexural cracks were visible to discussed in terms of crack opening rather than strain; the postpeak
the naked eye during the first two test stages, indicating extremely strain values of Region 3 in Fig. 6(b) reflect the observed longitu-
tight cracks that were efficiently controlled by the fiber and strand dinal extensions within the gauge length.
reinforcements. The detection of cracks was not investigated during The stress in the prestressing strands at the beginning of the
the third test stage due to safety concerns. flexural test, including all losses between strand release and time
The flexural capacity of the girder was achieved at an applied of test, was estimated by utilizing the concrete strains measured
moment of 7,271 kN-m (5,362 kip-ft) and a midspan deflection of by four vibrating wire gauges (VWGs) cast into the midspan cross
234 mm (9.21 in.). As the capacity was reached, a localized crack section. Two gauges were embedded between the top strands and
appeared at the midspan, initiating at the extreme bottom fiber and two at the centroid of the bottom strands. The initial concrete strain
propagating swiftly up through the bottom half of the web. The profile was then constructed by comparing the concrete strains
localization of the crack resulted in an abrupt decrease of approx- at the time of test to the reference strains taken prior to the deten-
imately 13.2% in the applied moment as shown in Fig. 6(a). A sioning of the strands. The difference in concrete strains at the level
photo of the crack pattern at the midspan immediately after locali- of each layer of strands was then determined assuming a perfect
zation is shown in Fig. 7(a) (crack pattern shown for this reading is bond between the strands and the surrounding concrete. When
indicated by a star in Fig. 6). After localization, the girder began these strains are multiplied by the modulus of elasticity of the
to hinge about the cross section at the localized crack. This behav- strands, Ep ¼ 196.5 GPa (28,500 ksi), the difference in the stresses
ior is unique to strain-hardening fiber-reinforced concrete beams, in the strands between jacking and the time of test can be estimated.
wherein: (1) the significant increase in curvature at the localized The effective prestress in the first (top), second, and third (bottom)
crack will contribute to a major portion of the overall deflection layers of strands at the midspan at the start of the test was calculated
of the beam; and (2) the curvature at the nonlocalized cracks will to be −1,336 MPa (−193.7 ksi), −1,157 MPa (−167.8 ksi), and
be controlled by the unloading and reloading stiffness of these sec- −1,144 MPa (−166.0 ksi), corresponding to strains of εp1;ini ¼
tions where the UHPC is exhibiting prelocalization behavior. The −0.00680, εp2;ini ¼ −0.00589, and εp3;ini ¼ −0.00582, respec-
large deformations at the localized crack cross section locally tively (tension taken as a negative value).
strained the bottom strands until they ruptured, separating the girder The initial stress profile in the girder at the beginning of the test
into two pieces only connected by the top two strands. A photo of was obtained from an elastic analysis of the cross section at the
the girder after strand rupture is shown in Fig. 7(b). After the midspan and using a concrete modulus of elasticity of 43.6 GPa
localization-induced applied moment reduction, the applied mo- (6,317 ksi), taken as the average of the two elastic modulus values
ment marginally increased with increasing displacement until the reported for each batch (Table 1). The initial stresses at the midspan
strands ruptured at an applied moment of 6,467 kN-m (4,769 kip-ft) were σT;ini ¼ 0.94 MPa (0.136 ksi) at the extreme top fiber
and midspan deflection of 287 mm (11.3 in.). (compression taken as a positive value) and σB;ini ¼ 29.7 MPa

east load point


west load point east load point
west load point

localized crack
localized crack

(a) (b)

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.

© ASCE 04022013-7 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


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
Parameter First crack 80% of yieldb Yieldc Localization Strand rupture
M exp kN-m (kip-ft) 4,122 (3,040) 4,884 (3,602) 6,983 (5,150) 7,775 (5,734) 6,970 (5,141)
εT a 0.00099 0.00123 0.00238 0.00299 0.00482d,e
εB a −0.00036 −0.00072 −0.00376 −0.00618 −0.03143d,e
εpb3 a −0.00672 −0.00705 −0.01000 −0.01239 −0.03611d,e
X mm (in.) 660 (26.0) 568 (23.4) 349 (13.8) 293 (11.5) 120 (4.72)e
ψ 1=mm (1=in:) 1.50 × 10−6 2.16 × 10−6 6.82 × 10−6 1.02 × 10−5 4.02 × 10−5
(3.80 × 10−5 ) (5.49 × 10−5 ) (1.73 × 10−4 ) (2.58 × 10−4 ) (1.02 × 10−3 )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., f p3 ¼ 0.80f̄ py , as defined in Table
5.9.2.2-1 of AASHTO LRFD BDS (AASHTO 2020).
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

c
Calculated when the stress in the last layer of strands is equal to the yielding stress limit of prestressing steel, i.e., fp3 ¼ f̄ 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.

(4.31 ksi) at the extreme bottom fiber, corresponding to strains of with conventional steel and/or prestressing strands. The framework
εT;ini ¼ 0.000022 and εB;ini ¼ 0.000682, respectively. is founded on structural mechanics principles commonly used in
A summary of the test results at the midspan, including the ef- conventional reinforced concrete design where the nominal flexural
fect of the prestressing force and considering the self-weight of the capacity at each cross section of the beam is calculated by satisfy-
girder and loading apparatus, is calculated at the five points in ing the conditions of equilibrium and strain compatibility. Because
flexural behavior, namely, the initiation of cracks; service stress the mechanical behavior of UHPC is different from that of conven-
limit of prestressing steel taken when the stress in the bottom layer tional concrete, new constitutive relationships are proposed and
of strands is equal to 80% of the strand yield stress (AASHTO then utilized within the strain compatibility approach to predict the
2020), i.e., f p3 ¼ 0.80f̄py ; yielding of the bottom layer of strands, complete nominal moment–sectional curvature (M n -ψ) diagram
i.e., fp3 ¼ f̄py ; crack localization; and strand rupture as presented and the flexural failure modes of UHPC beams. In this framework,
in Table 2. In Table 2, M exp is total moment corresponding to ap- it is assumed that (1) sections plane before bending remain plane
plied moment M plus the self-weight moment M ini , as described in after bending; (2) the member is reinforced with conventional steel
Eq. (1); εT and εB are the total concrete strains in the top and bottom and/or prestressing strands within the flexural tension side; (3) a
layers calculated according to Eqs. (2) and (3), in which ΔεT and perfect bond exists between the reinforcement and UHPC, i.e., the
ΔεB are the change in strains captured in extreme top and bottom change in strain in the bonded reinforcement and/or prestressing
fibers during the test within Region 3; εp3 is the strain in the bottom strands is equal to the strain in the surrounding UHPC; (4) the
layer of strands calculated according to Eq. (4), in which εp3;ini ¼ stresses in the reinforcing steel or prestressing strands are derived
−0.00582 is the initial strain in the bottom layer of strands and from the uniaxial stress-strain response indicative of their tensile
Δεcp3 is the change in strain in the concrete at the level of the bot- behavior; (5) the composition of UHPC mixture and its mechanical
tom layer of strands during the test within Region 3; X is the neutral properties meet specified threshold values; and (6) the stresses
axis depth calculated from the strain profile formed by εT and εB ; in the UHPC are inferred from appropriate uniaxial stress-strain
and ψ ¼ εT =X is the sectional curvature at the midspan. The 80% models. The specified threshold values and the uniaxial stress-
of yield and yield moment values reported in Table 2 and Fig. 6 strain models for UHPC are discussed in the following section.
were determined when εp3 reached a strain equal to -0.00705 (cor-
responding to 0.80f py ) and the yielding strain of the strands,
εpy ¼ −0.01, respectively UHPC-Class Materials and Mechanical Models
A key starting point for the proposed flexural design framework is
Mexp ¼ M þ M ini ð1Þ the delineation of the scope of the considered UHPC-class materi-
als and the definition of the fundamental parameters. UHPC is gen-
εT ¼ εT;ini þ ΔεT ð2Þ erally defined as portland cement composite made of an optimized
gradation of granular constituents, a water-to-cementitious materi-
εB ¼ εB;ini þ ΔεB ð3Þ als ratio less than 0.25, and a high percentage of discontinuous in-
ternal steel fiber reinforcement. The specific mechanical properties
εp3 ¼ εp3;ini þ Δεcp3 ð4Þ 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
Finally, the slippage of selected strands during the test was to sustain the cracking strength through a minimum localization
monitored with three LVDTs mounted on three of the bottom-layer strain of at least 0.0025.
strands located within one half of the section on the west end of the The compressive properties, namely, the modulus of elasticity
girder. As expected with a beam of this length, these measurements Ec , the compressive strength fc0 , and the strain at compressive
demonstrated that no strand slippage occurred during the test. strength εcu , are obtained from cylindrical specimens tested accord-
ing to ASTM C39 (ASTM 2018b) and ASTM C469 (ASTM 2014)
with provisions specific to UHPC described in ASTM C1856
Description of Proposed Flexural Design Method for (ASTM 2017). The uniaxial elastic-perfectly plastic model pro-
UHPC Beams posed by El-Helou et al. (2022) is adopted in this framework to
determine the stress in the UHPC as a function of the compression
The proposed flexural design framework described in this section is strain in the section, as shown in Fig. 8(a). This model mimics the
relevant to the design of beams made with UHPC and reinforced experimental uniaxial stress-strain response in the elastic region up

© ASCE 04022013-8 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


Stress

Stress

Stress
fc’

ft,loc

ft,cr ft,cr

Ec
Ec Ec

cp cu Strain t,cr t,loc Strain t,cr t,loc Strain


(a) (b) (c)
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

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.

to a reduced compressive stress of αf c0 , where α is a reduction fac- Flexural Behavior of UHPC Beams
tor on compressive strength reflecting the linearity limit of the com-
The flexural behavior of UHPC beams can be obtained analytically
pressive stress-strain response. After this stress level, the model
by employing a strain compatibility approach and utilizing the con-
sustains the reduced compressive resistance until the strain at
stitutive models for UHPC and steel reinforcement depicted in
the material’s compressive strength εcu is reached. El-Helou et al.
Figs. 8 and 9, respectively. For flexural members fully utilizing
(2022) suggested a reduction factor α value of 0.85, based on the
the UHPC tensile resistance, the nominal moment–sectional curva-
experimental results of five commercially available UHPC products
ture (M n − ψ) diagram can be idealized in four key points, as illus-
(Graybeal and Stone 2012; Haber et al. 2018).
The tension parameters, namely, the effective cracking strength trated in Fig. 10, for a UHPC member with conventional steel
f t;cr , the localization stress ft;loc , and the localization strain εt;loc , reinforcement. The sectional curvature is defined as the ratio of
are obtained from prismatic specimens tested in direct tension ac- the strain in the extreme compression layer εc divided by the depth
cording to the method developed by Graybeal and Baby (2013). of the neutral axis X: ψ ¼ εc =X. The first key point in the behavior
This method delivers the uniaxial stress-strain responses of the (ψcr , M cr ) represents the initiation of the first flexural tensile crack,
tested specimens, facilitating the identification of critical tension occurring when the strain in the extreme tension layer reaches the
parameters. The effective cracking strength ft;cr , is taken as the cracking strain limit of UHPC, εt;cr ¼ f t;cr =Ec . At this point, the
intercept of a line having a slope equal to the elastic modulus compression and tensile stresses remain within the elastic region of
and a strain offset of 0.02%. The localization stress ft;loc and strain their respective constitutive models (Fig. 8). The second point in
εt;loc are taken as the stress and strain of the data point where the the moment-curvature diagram (ψy , My ) coincides with the yield-
stress begins to continuously decrease with increasing strain. As ing of the extreme tension layer of the reinforcing steel, occurring
proposed by El-Helou et al. (2022), the constitutive law for UHPC when the steel strain is equal to the yielding strain of conventional
in tension idealizes the experimental curve into an elastic-perfectly steel εsy or prestressing strands εpy . Note that in Fig. 10, an inter-
plastic relationship, as shown in Fig. 8(b), for material exhibiting a mediate point (ψsl , M sl ) in behavior is shown corresponding to the
stress plateau after cracking, or a bilinear relationship, as shown in chosen baseline sectional curvature and moment for ductility con-
Fig. 8(c), for material exhibiting a continuous increase in stress siderations as discussed later in the manuscript. The third key point
after cracking and when the localization stress f t;loc is at least (ψL , M L ) indicates the flexural nominal capacity of the section
20% greater than the effective cracking stress f t;cr . El-Helou et al. (M n ¼ ML ) and coincides with the crack localization, where the
(2022) proposed that a reduction factor γ be applied on the tensile tensile capacity of UHPC is fully utilized and the strain in the ex-
stress parameters (ft;cr and f t;loc ) to account for variability in the treme tension layer is equal to the localization strain εt;loc . The
material behaviors; the value of the reduction factor γ is not to cross-sectional strain and stress diagrams at the flexural crack
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; f sy
Stress

Stress

fpu
and εsy are the yielding stress and strain, respectively; and εsu is the
strain at rupture of the bars. The idealized stress-strain relationships fpy
for prestressing strands can be obtained from existing nonlinear
models, such as the power formula based on the work of Skogman fsy
et al. (1988) shown in Fig. 8(b). For instance, the stress in the pre-
stressing steel fps for 1,860 MPa (270 ksi) seven-wire low- Ep
relaxation strands can be computed as a function of the steel strain Es
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 sy Strain su py Strain pu
27,613 ksi, and 1,860 MPa with 270 ksi (PCI 2014) (a) (b)
 
190,385 Fig. 9. Idealized uniaxial stress-strain relationships for (a) conventional
fps ¼ εps 6,116 þ 1 ≤ 1,860 MPa ð5Þ steel reinforcing bars; and (b) prestressing strands.
ð1 þ ð112.4εps Þ7.36 Þ7.36

© ASCE 04022013-9 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


Mn Beam Strains Stresses
c = cu fc = f’c
ML
compression
My M X
Mc neutral
axis ft,cr
d t,cr
h tension
Msl
t,loc ft,cr
Mcr
s fs
t > t,loc

cr sl y L c Fig. 12. Cross-sectional strain and stresses conditions for UHPC mem-
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

bers in flexure at compression failure and after the localization of


Fig. 10. Idealized nominal moment versus sectional curvature (M n −
cracks (shown for a UHPC exhibiting a stress plateau after cracking
ψ) diagram of UHPC member with conventional steel reinforcement.
in tension).

localization point for a UHPC material exhibiting a stress plateau


remains less than the strain at rupture of the bars εsu when the com-
after cracking [Fig. 8(b)] are shown in Fig. 11, where the compres-
pressed UHPC crushes. However, due to the high compressive
sion stresses remain elastic with the strain value in the extreme
strength of UHPC materials, the rupture of the steel reinforcement
compression layer εc, less than εcp ¼ αfc0 =Ec [Fig. 11(a)], or
(conventional steel or prestressing strands) can occur before con-
become plastic when εcp < εc ≤ εcu [Fig. 10(b)]. In Fig. 10, it is
crete crushing (i.e., when the strain in the extreme compression
assumed that the localization strain of UHPC εt;loc is greater than
layer is less than εcu ). In an effort to engage the full compressive
the yielding strain of the reinforcement εsy , and thus the yielding
strength of UHPC in a flexural design, Shao and Billington (2019)
point (ψy , M y ) occurs at a smaller curvature and moment values
than those defining the localization point (ψL , M L ). For prestressed demonstrated that the postlocalization moment capacity of UHPC
members, the yielding point occurs before the localization point if beams can be carried through high-tensile strains until compression
the change in the prestressing steel strain between the effective pre- crushing (Mc ≥ ML ) when the beam is heavily reinforced with con-
stress and yield (εpy − εpe ) is greater than the UHPC localization ventional steel reinforcement and when the strain-hardening char-
strain εt;loc . After crack localization (εt > εt;loc ), the UHPC can no acteristics of the steel are both engaged and substantial. While such
longer contribute to the tension resistance of the member, leading to behavior is desirable because it significantly increases the ductility
a reduction in the flexural capacity. The last point in the flexural of UHPC beams, to date the approach has only received limited
behavior shown in Fig. 10 (ψc , M c ) coincides with the compression experimental investigation. When the tensile strain in the UHPC
failure of UHPC, occurring when the strain in the extreme compres- exceeds the localization strain at a crack (εt > εt;loc ), the loss of
sion layer reaches the ultimate compressive strain of UHPC εcu . the UHPC fiber-bridging resistance mechanism results in a change
The cross-sectional strain and stress diagrams at compression fail- in the internal force-resisting mechanism (i.e., increased curvature,
ure and after the localization of cracks (εt > εt;loc ) are depicted in increased strains, and movement of the neutral axis). If the in-
Fig. 12 for a UHPC material exhibiting a stress plateau after crack- creased tensile strain results in increased tensile resistance due to
ing [Fig. 8(b)]. The shape of the postlocalization moment-curvature the engagement of additional discrete reinforcement force-resisting
relationship is nonlinear (dashed line in Fig. 10) and depends on the mechanisms, such as the strain-hardening of steel or the further en-
UHPC material properties and the design parameters of the section. gagement of unbonded reinforcements, then a new stable moment-
Intermediate points between localization and compression failure resisting state may be realized. For this reason, postlocalization
need to be computed if the exact shape of the postlocalization curve analysis should take into consideration the width of the localized
is desired. crack, the local straining of the reinforcement crossing the crack,
In the determination of the moment-curvature diagram in and the resistance provided by the reinforcement. In the current
Fig. 10, it is assumed that the strain in the steel reinforcement εs proposal, the postlocalization flexural capacity is not recommended

Beam Strains Stresses Strains Stresses


c cp fc f’c cp c cu fc = f’c

compression X X
M
neutral
d axis
h ft,cr ft,cr
tension t,cr t,cr

s
fs s fs
ft,cr ft,cr
t = t,loc t = t,loc
(a) (b)

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).

© ASCE 04022013-10 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


for use in design because more research is needed to quantify namely, the effective cracking stress f t;cr and the localization stress
the relationship between the rupture strain of the bars after UHPC ft;loc and strain εt;loc , can be obtained by executing direct tension
crack localization, the opening of the localized crack, and the de- tests proposed by Graybeal and Baby (2013).
velopment length of the reinforcement in the vicinity of the local-
ized crack. The development length of the reinforcement is a Service and Ultimate Limit States
critical consideration as it significantly influences the length of The flexural analysis can be performed using a strain compatibility
the bar or strand over which the increased tensile strain demand approach where stresses over strained cross sections are calculated
will be resisted. and then summed to satisfy equilibrium requirements. In a situation
For flexural members not fully utilizing the UHPC tensile resis- where the cracking of the girder is not allowed at service, the lin-
tance, the ultimate capacity occurs at compression crushing (M n ¼ early elastic analysis procedures typically used with conventional
M c ), when the UHPC in the extreme compression layer reaches its concrete design can be employed, with a tensile stress limit taken as
ultimate compressive strain limit (εc ¼ εcu ) while the strain in the a fraction of the effective cracking stress, e.g., f t;lim ¼ 0.85ft;cr .
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

extreme tensile fiber is less than the localization strain (εt < εt;loc ). When a minimal amount of cracking is permissible at service, a
In these situations, the beam design does not take full advantage of strain compatibility analysis can be performed in which the strains
the UHPC tensile performance characteristics, prompting a com- in the tensile zone are limited to a fraction of the localization strain,
pression failure at reduced tensile strain values, such as could occur e.g., εt;lim ¼ 0.30εt;loc , and the stresses in the reinforcing steel are
in sections having a very high reinforcement ratio and/or sections lower than a fraction of the yielding stress, e.g., fs;lim < 0.80f sy and
subjected to high axial compressive loads. The nominal flexural fp;lim < 0.80fpy . In both service limit state cases, the compression
moment–sectional curvature (M n − ψ) for this failure mode can stress limits outlined in AASHTO LRFD BDS (AASHTO 2020)
be defined by straight lines, with either (1) three key points coincid- for conventional concrete can also be used for UHPC. For the ul-
ing with the initiation of first flexural tensile crack (ψcr , Mcr ), timate limit state, the stress-strain constitutive models in Fig. 8 are
yielding of the extreme tension layer of the reinforcing steel used, allowing the construction of the full moment-curvature dia-
(ψy , M y ), and the compression crushing of UHPC (ψc , Mc ) when gram. In this case, the linear strain distribution along the height of
the steel strain at crushing is greater than the yield strain; or (2) two the section is bounded by the UHPC crushing strain limit εcu , the
key points coinciding with the initiation of cracks (ψcr , M cr ) and localization strain εt;loc , and the rupture strain limit of conventional
compression crushing (ψc , Mc ) when the steel strain at crushing is steel εsu or prestressing strands εpu . The nominal moment capacity
less than or equal to the yield strain. is defined at either (1) crack localization (M n ¼ M L ), when the
strain in the extreme tensile layer εt is equal to εt;loc , which can
occur when the compression stress is elastic (εc ≤ εcp ) or plastic
Proposed Design Methodology (εcp < εc ≤ εcu ), as shown in Fig. 11; or (2) compression failure
(M n ¼ Mc ) when the strain in the extreme compression layer εc
Based on the experimentally observed behaviors and analytical re-
is equal to εcu and the strain in the extreme tensile layer εt is less
sults, a flexural design framework for UHPC beams reinforced with
than εt;loc .
conventional steel or prestressing strands is proposed in this section.
The design framework is similar to portions of existing UHPC struc- Consideration of Curvature Ductility
tural design guidance detailed in various international documents In cases where the flexural failure is governed by the localization of
[NF P18-710 (AFNOR 2016a); SIA 2052 (SIA 2016); CSA the cracks (M n ¼ M L ), it is important that members exhibit ad-
S6:19 (CSA 2019)] and is tailored to generally parallel existing equate ductility to ensure that large deformations occur before fail-
bridge design specifications in the United States (AASHTO 2020) ure of the member. An approach based on sectional curvature
while accounting for UHPC mechanical performance. ductility can be adopted in which a curvature ductility ratio μ is
The first step in the structural design is the characterization of defined and compared to a specified minimum curvature ductility
the material-level properties and the establishment of constitutive ratio, e.g., μmin ¼ 3.0. The curvature ductility ratio can be taken as
laws that describe the mechanical response when subjected to struc- the ratio of the sectional curvature at localization ψL and a baseline
tural loads. For UHPC in compression, the parabolic stress-strain sectional curvature, ψsl : μ ¼ ψL =ψsl . It is proposed that the base-
relationship typically used for conventional concretes is replaced line sectional curvature ψsl be calculated when the stress in the ex-
with the idealized elastic-perfectly plastic model in Fig. 8(a). treme tension layer of reinforcement is equal to 80% of the yielding
The compression model requires the modulus of elasticity Ec, com- stress of the steel reinforcement, i.e., 0.80f py or 0.80f sy. For
pressive strength f c0 , and strain at compressive strength εcu as input prestressed members, this limit corresponds to maximum stress that
parameters, which can be experimentally obtained from compres- can be attained in the prestressing steel at the service limit state
sion tests performed according to ASTM C1856 (ASTM 2017). specified in AASHTO LRFD BDS (AASHTO 2020). When
In lieu of experimental testing, the elastic modulus can be approxi- members have a curvature ductility ratio less than μmin , a reduced
mated by Eq. (6) and εcu can be taken as the greater of εcp ¼ resistance factor on the nominal moment M n is recommended,
αf c0 =Ec and 0.0035, as proposed by El-Helou et al. (2022). in recognition of the nonductile failure mechanism akin to
The reduction factor α can be taken as 0.85 compression-controlled failure. The proposed resistance factors
are discussed later in the manuscript.
Ec ¼ 9,100fc00.33 ðMPaÞ Limited research on reinforced UHPC members has shown that
the localization strain value is increased when the cracking resis-
Ec ¼ 2,500fc00.33 ðksiÞ ð6Þ
tance provided by the fibers is supplemented by conventional steel
rebars or prestressing strands (Oesterlee 2010; Baby et al. 2014).
In tension, UHPC offers a postcracking resistance that is sus- For instance, the Swiss recommendation allows the use of a
tained through large tensile strains (εt > 0.0025), delivering a ben- localization strain in the tensile mechanical model that is twice
eficial contribution to the flexural capacity of a beam. The tensile the value obtained from material testing (SIA 2016). In the current
constitutive models for UHPC can be idealized with the elastic- study, the measured strain at localization of the flexural crack
perfectly plastic or bilinear (when f t;loc ≥ 1.20f t;cr ) relationships (Table 2) was approximately 24% greater than the localization
shown in Figs. 8(b) and 7(c), respectively. The tensile parameters, strain obtained from a direct tension test (Batch A in Table 1). The

© ASCE 04022013-11 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


implementation of a localization strain value higher than the one tension-controlled sections for conventional concrete members
obtained from tension testing in flexural design would allow for in AASHTO LRFD BDS (AASHTO 2020)
the use of a larger tensile strain, thus increasing the likelihood of
satisfying the ductility requirement for many UHPC beams. How- 0.15ðμ − 1.0Þ
ϕf ¼ 0.75 þ ; 0.75 ≤ ϕf ≤ 0.90 ð8Þ
ever, more research is needed before broad implementation of μmin − 1.0
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 dis- Validation of the Proposed Flexural Model for
crete tensile reinforcement in the flexural element. Finally, as pre- UHPC Beams
viously discussed, the true ductility of a member may be greater
than that analytically predicted using the aforementioned pro- The capability of the model to predict the flexural behavior of pre-
cedure if the opening of a localization crack generates sufficient stressed and non-prestressed UHPC beams was evaluated by com-
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

increased resistance from the tensile reinforcement (e.g., through paring the model predictions to the experimental data from 45
engagement of strain-hardening behaviors in mild steel reinforce- flexural tests performed by the authors or found in the literature.
ment) that the moment at compression crushing M c becomes
greater than M L . This topic is the subject of future research by
the authors. Prestressed Girder Test Presented in this Paper
The predicted moment-curvature behavior of the prestressed
Consideration of Fiber Orientation
girder tested by the authors, and previously discussed in this
The orientation of the fibers within the tensile zone of beams has a
paper, was obtained by discretizing its cross section into
significant effect on the tensile response of UHPC. Fibers tend to
2.5-mm-(0.1-in.)-thick horizontal layers through the height and
preferentially align with the direction of flow of fresh UHPC during
assuming constant strains within each layer. The initial strains in
casting, with dissimilar tensile stress and strain capacities being ex-
each layer at the beginning of the test were determined from
hibited between the flow and nonflow directions (Maya Duque
the UHPC strain measured by the VWGs embedded at the midspan
et al. 2016; Doyon-Barbant and Charron 2018; Huang et al. 2018;
and following the analytical considerations described previously.
Walsh et al. 2018; Qiu et al. 2020b). For this reason, it is recom-
That is, the initial strain in the first (top), second, and third (bottom)
mended that a casting method be specified by the designer to ensure
layer of strands is equal to εp1;ini ¼ −0.00680, εp2;ini ¼ −0.00589,
that the fiber orientation in the in-place material is similar to the one
and εp3;ini ¼ −0.00582, respectively (tension taken as a negative
achieved in the material control specimen. Unfavorable fiber ori-
value). The initial strain in each layer of UHPC was calculated from
entation with respect to the flexural tensile strains can reduce the
the initial strain profile defined by the top strain, εT;ini ¼ 0.000022,
expected cracking stress and localization strain capacities of the
and bottom strain, εB;ini ¼ 0.000682 (compression taken as a pos-
UHPC in the structure. In these cases, a reduction factor γ lower
itive value). The total strain in the UHPC section was obtained by
than the maximum value of 0.85 is recommended to be applied to
superposing the initial strain profile with an assumed strain profile,
the stress and strain parameters of the tensile constitutive law, as
i.e., εi ¼ εi;ini þ Δεi , in which εi , εi;ini , and Δεi are the total, initial,
shown in Figs. 8(b and c), to ensure a conservative representation
and assumed strains in the UHPC layer i, as shown in Fig. 13. Sim-
of the tensile response in the structure. Determining the value of γ
ilarly, the total strains in the strand layers, εp1 , εp2 , and εp3 , were
to address the fiber orientation effects may, in some instances, re-
calculated by superposing the initial steel strains, εp1;ini , εp2;ini , and
quire prototype testing as recommended by the French design stan-
εp3;ini , with the change in strain in the UHPC at the level of each
dard NF P18-710 (AFNOR 2016a).
steel layer, Δεcp1 , Δεcp2 , and Δεcp3 , respectively. The stresses in
each layer were calculated by employing the mechanical models
Factored Flexural Resistance
shown in Figs. 8(a and b) and 9(b). The stresses were then con-
At present, there is insufficient data to allow for the calibration of
verted into compression and tension forces, which were summed
the resistance factor for UHPC beams. However, the LRFD frame-
to check force equilibrium in the section. This process was repeated
work requires that a factor be applied to the resistance calculation.
by assuming different strain profiles until equilibrium was satisfied,
Thus, it is proposed that the factored flexural resistance M r of a
after which the flexural moment was computed. The procedure was
UHPC beam can be taken in accordance with the relationship
repeated by increasing the values of assumed strains until the full
shown in Eq. (7), with ϕf being the resistance factor for flexural
moment–sectional curvature diagram was derived.
moment
In calculating the UHPC stresses, the compression and tension
parameters were informed from the average results of the material
M r ¼ ϕf M n ð7Þ tests performed on companion specimens presented in Table 1. The
UHPC modulus of elasticity was taken as the average of the results
For members dominated by compression crushing (i.e., M n ¼ M c obtained from the two UHPC batches with Ec ¼ 43.6 GPa (6,317
with εc ¼ εcu and εt < εt;loc ), the resistance factor can be taken Ksi). The compressive strength and ultimate compression strain
equal to 0.75. This value was chosen to correspond to the resis- were taken from the results of Batch B, which filled the top half
tance factor specified for compression-controlled sections in of the girder: fc0 ¼ 161 MPa (23.3 ksi) and εcu ¼ 0.00401. The
AASHTO LRFD BDS (AASHTO 2020). For members dominated tensile parameters were taken from the results of Batch A, which
by crack localization (i.e., M n ¼ M L with εc ≤ εcu and εt ¼ εt;loc ), filled the bottom half of the girder, and without a reduction factor
the resistance factor can be taken as 0.90 for sections satisfying (i.e., γ ¼ 1.00): γf t;cr ¼ γft;loc ¼ −9.3 MPa (−1.35 ksi) [elastic-
the ductility requirement (μ ≥ μmin ) and 0.75 for nonductile perfectly plastic model in Fig. 8(b)] and εt;loc ¼ −0.00497. The
sections with a sectional ductility ratio less than or equal to tensile design model is compared to the uniaxial tensile test results
1.0. For prestressed and non-prestressed members falling in be- in Fig. 5(a). The steel stress-strain model relationship for the
tween the two extremes, a linear variation of the resistance factor 1,860 MPa (270 ksi) low-relaxation strands was determined ac-
is proposed in Eq. (8). This transition zone is proposed akin cording to Eq. (5) and depicted atop the experimental data from
of the transitional zone between compression-controlled and the strand pullout tests in Fig. 5(b). For strand stress values greater

© ASCE 04022013-12 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


Girder Cross-Section Initial Strains Change in Strains Total Strains
T,ini T T
cp1,ini cp1 cp1
X
X

cp2,ini cp2 cp2


cp3,ini cp3 cp3
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

B,ini B B

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.

than 1,860 MPa (270 ksi) or strain values greater than 0.0272, a The experimental and predicted design responses in Fig. 14(a)
linear interpolation was implemented up to a rupture stress fpu are in good agreement, demonstrating the capability of the pro-
of 1,932 MPa (281 ksi) occurring at a strain εpu of −0.0674, as posed method to predict the full moment–sectional curvature re-
shown in Fig. 5(b). Table 3 presents a summary of the strain com- sponse of prestressed beams. The results also confirm that the
patibility analysis results at the five points in flexural behavior, shape of the moment–sectional curvature diagram can be idealized
namely, the initiation of cracks; strand service stress limit, i. by four key points, as illustrated in Fig. 10. Because the compres-
e., 0.80fpy , reached in the bottom layer of strands; yielding of sion strain limit observed during the test was approximately 20%
the bottom layer of strands; crack localization; and compression larger than the ultimate compression strain specified in the simu-
crushing. These results represent the nominal design capacities lations, the predicted failure mode of the girder was compression
of the girder and can be directly compared to the experimental data crushing instead of strand rupture, as observed in the test [Table 3
summarized in Table 2. The predicted-over-experimental flexural and Fig. 14(a)]. This result suggests that the UHPC in a structural
moment ratio, Mn =M exp , is 0.97, 1.09, 0.99, and 0.92 at first crack, element can undergo compression strains exceeding the strain at
80% of yield, yield, and localization, demonstrating the capability peak stress, as obtained from a uniaxial compression test, likely
of the proposed methodology to predict the flexural capacity of pre- due to the confinement effects within the larger beam and strain
stressed UHPC beams. The predicted nominal design curve is com- gradient effect within a beam cross section with large curvature.
pared to the experimental results in Fig. 14, plotted as the Similarly, the localization strain of UHPC observed during the test
relationship between the applied moment and the change in curva- (εB ¼ −0.00618 in Table 2 at localization point) was approxi-
ture during the test and simulation. The simulated design and mately 24% greater than the localization strain obtained from direct
experimental moments in Fig. 14 are taken as M n − M ini and tension testing [εB ¼ ε̄t;loc ¼ −0.00497 in Table 1 (Batch A) and
M exp − M ini , respectively, where Mini ¼ 467 kN − m (345 kip-ft) Table 3 at localization point], indicating that the primary tensile
is the initial moment in the beam at the beginning of the test. reinforcement and strain gradient effects may enhance the flexural
The change in curvature Δψ is taken as ΔεT =ΔX, where ΔX is tensile strain capacity of the UHPC.
the depth of the neutral axis of the intermediate strain diagram The design curvature ductility ratio of the girder, μsim ¼ 3.50,
shown in Fig. 13, a value calculated in the strain compatibility can be calculated by dividing the sectional curvature at localization
analysis and measured during the test. Note that the simulated over the sectional curvature at 80% of yield reported in Table 3.
design curve was obtained from 110 calculation points, which is Because the ductility ratio is greater than the specified ratio μmin
greater than the minimum four key points proposed in the design of 3.0, the flexural resistance factor for this girder is equal to 0.90.
methodology. Note that the experimental ductility ratio obtained from the results

Table 3. Summary of the predicted results at midspan computed according to the proposed flexural framework
Parameter First crack 80% of yieldb Yieldc Localization Crushingd
M n kN-m (kip-ft) 4,015 (2,961) 5,306 (3,913) 6,921 (5,105) 7,132 (5,260) 6,636 (4,894)
εT a 0.00099 0.00139 0.00226 0.00245 0.00401
εB a −0.00021 −0.00073 −0.00391 −0.00497 −0.02262
εpb3 a −0.00661 −0.00708 −0.01000 −0.01102 −0.02759
X mm (in.) 741 (29.2) 590 (23.2) 330 (13.0) 297 (11.7) 136 (5.3)
ψ1=mm (1=in:) 1.34 × 10−6 2.35 × 10−6 6.85 × 10−6 8.23 × 10−6 2.96 × 10−5
(3.40 × 10−5 ) (5.97 × 10−5 ) (1.74 × 10−4 ) (2.09 × 10−4 ) (7.51 × 10−4 )
M n =Mexp 0.97 1.09 0.99 0.92 0.95
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., f p3 ¼ 0.80f̄ 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 ¼ f̄ py .
d
Calculations assume strain compatibility is maintained after localization.

© ASCE 04022013-13 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


Sectional Curvature (in-1) 10-3 Sectional Curvature (in-1) 10-3
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
8,000
5,500

Applied Moment at Midspan (kip-ft)


Applied Moment at Midspan (kN-m)
7,000 Nominal design curve
5,000
6,000 4,500
Fitted curve
4,000
5,000 Factored design curve (pre-localization)
Experimental curve Experimental curve 3,500
4,000 UHPC tensile 3,000
behavior ignored 2,500
3,000
First crack 2,000
80% of yield stress in bottom layer of strands
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

2,000 1,500
Yielding of bottom layer of strands
Localized crack (ultimate moment) 1,000
1,000 Rupture of bottom strands
500
Compression crushing
0 0
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
Sectional Curvature (mm-1) 10-5 Sectional Curvature (mm-1) 10-5
(a) (b)

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.

reported in Table 2 is greater than μsim with μexp ¼ 4.72 because (rectangular shapes for non-prestressed, and I and pi shapes for
the cracks in the girder localized at a higher localization strain than prestressed). The height of the non-prestressed rectangular beams
the one used in the simulations. The factored moment-curvature ranged between 220 mm (8.7 in.) and 381 mm (15 in.), while the
design curve is compared to the nominal and experimental curves height of the prestressed I- and pi-shaped beams were 838 mm
in Fig. 14(a). (33 in.) and 914 mm (36 in.). The beams were made with different
To illustrate the key differences between the behavior of pre- types of UHPC products with varying amounts of fiber volume
stressed conventional concrete beams and similar UHPC beams, content (1.0%, 1.5%, and 2%) and prestressing or conventional
a strain compatibility analysis of the same girder was performed steel reinforcement (reinforcement ratio ρ ranging between 0%
ignoring the tensile resistance of UHPC, i.e., ft;cr ¼ f t;loc ¼ 0, and 4.99%). The fibers varied in shape and geometries, including
with the results plotted in Fig. 14(a). While this analysis resulted smooth, twisted, straight, and hooked-end, with diameters and
in a nominal moment capacity (M n ¼ Mc ) lower than the experi- lengths ranging between 0.12 and 0.30 mm (0.0047 and 0.012 in.)
mental capacity, it both incorrectly predicted a concrete crushing and 10.0 and 30.0 mm (0.39 and 1.18 in.), respectively. The re-
failure mode at the top of the section, rather than localization of ported flexural behaviors of the beams are in line with the flexural
cracks at the bottom, and also significantly overestimated the cur- behavior of UHPC members described previously in this paper: the
vature ductility ratio μnt;sim compared to the experimental ductility maximum flexural moment was reached immediately before a sin-
ratio, i.e., μnt;sim ¼ 2.67μexp . This comparison demonstrates that gle localized crack—initiated by the pullout of the crack-bridging
conducting a capacity-based design that ignores the UHPC tensile fibers—occurred, followed by the subsequent rupture of the tensile
resistance may inadvertently result in a structure that does not dem- reinforcement. Therefore, the analytical predictions presented in
onstrate the desired ductility. this section focus on determining the capacity of the beams at
Finally, if the localization strain measured during the girder test, the onset of localization, as recommended in the proposed design
εt;loc ¼ −0.00636, is used in the model, the simulated nominal methodology that is developed in this paper. The predicted results
design curve is identical to the prelocalization experimental for the beams found in the literature are presented in Table 4. A
moment-curvature curve, as shown in Fig. 14(b), when the bilinear portion of the input material parameters for the UHPC and reinforc-
constitutive law in Fig. 8(c) is employed with the following cali- ing steel given in Table 4 were obtained from independent material
brated input values for the cracking and localization stresses: tests that accompanied each beam, i.e., simulations of beams tested
f t;cr ¼ −6.21 MPa (0.90 ksi) and ft;loc ¼ −15.5 MPa (−2.25 ksi). by Chen et al. (2018), Yoo et al. (2017), Qiu et al. (2020a), Yoo and
In this analysis, the estimated value of the cracking stress in the Yoon (2015), Graybeal (2006b), and Graybeal (2009). The uniaxial
girder was approximately 33% lower than the one obtained from tensile parameters (γf t;cr , γft;loc , and εt;loc with γ ¼ 1.00) for the
direct tension testing, while the localization stress was 49% higher UHPC in these beams were obtained from uniaxial testing methods
than the localization stress value obtained from tension testing or inverse analysis techniques performed on the results of flexural
(Batch A in Table 1). beam tests. In cases where key material parameters were not re-
ported, they were obtained by fitting the model results to the exper-
imental data from one of the tested beams in a set, then comparing
Prestressed and Non-Prestressed Beams Tested by the predicted flexural capacities of the remaining beams within the
Other Researchers
same set to the experimental results without changing the estimated
This section explores the experimental results from tests of two ad- input parameters, i.e., simulations of beams tested by Yang et al.
ditional prestressed girders and 42 non-prestressed reinforced (2010), Meade and Graybeal (2010), and Hasgul et al. (2018), as
beams found in the literature and compares them to model predic- indicated in Table 4. The predicted flexural moment capacity M n
tions. The beams had varying cross-sectional shapes and geometries of the beams presented in Table 4 is in good agreement with an

© ASCE 04022013-14 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


Table 4. Comparison between experimental and predicted flexural capacities for prestressed and non-prestressed UHPC beams found in the literature
Original f c0 E f t;cr ft;loc ρ f sy Mn M exp
Reference name (MPa) (MPa) εcu (MPa) (MPa) εt;loc (%) (MPa) (kN-m) (kN-m) M n =M exp
f g h h h
Chen et al. B-1 126 45.6 0.0035 9.8 11.9 0.0043 1.09 461 54.0 43.3 1.25
(2018)a B-2 126 45.6f 0.0035g 9.8h 11.9h 0.0043h 2.75 417 78.5 71.4 1.10
B-3 126 45.6f 0.0035g 9.8h 11.9h 0.0043h 3.60 456 96.2 90.4 1.06
B-4 126 45.6f 0.0035g 9.8h 11.9h 0.0043h 4.99 445 111.2 105.9 1.05
Yoo et al. UH-N 197 47.8 0.0044 11.1i 12.2i 0.0032i 0.00 N=A 66.5 72.5 0.92
(2017)a UH-0.53% 197 47.8 0.0044 11.1i 12.2i 0.0032i 0.53 523 91.9 97.9 0.94
UH-1.06% 197 47.8 0.0044 11.1i 12.2i 0.0032i 1.06 523 117.0 118.8 0.99
UH-1.71% 197 47.8 0.0044 11.1i 12.2i 0.0032i 1.71 523 128.1 131.0 0.98
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

Qiu et al. B-S65-16 132 44.0 0.0035g 8.0h 9.4h 0.0048h 0.73 412 83.8 65.5 1.28
(2020a)b B-S81-20 126 43.1 0.0035g 8.1h 9.7h 0.0062h 1.14 460 114.0 105.0 1.09
B-S83-20 130 45.3 0.0035g 8.0h 9.6h 0.0053h 1.14 460 113.0 101.9 1.11
B-H65-20 139 45.2 0.0035g 8.0h 10.4h 0.0067h 1.14 502 123.0 133.0 0.93
Yoo and Yoon S13-0.94% 212 46.7 0.0045 7.0i 7.0i 0.0023i 0.94 494 33.5 39.3 0.85
(2015)c S13-1.50% 212 46.7 0.0045 7.0i 7.0i 0.0023i 1.50 503 40.2 55.8 0.72
S19.5-0.94% 210 46.9 0.0048 10.6i 10.6i 0.0031i 0.94 494 44.8 42.0 1.07
S19.5-1.50% 210 46.9 0.0048 10.6i 10.6i 0.0031i 1.50 503 53.7 56.3 0.95
S30-0.94% 210 46.8 0.0046 9.8i 11.0i 0.0073i 0.94 494 47.0 43.2 1.09
S30-1.50% 210 46.8 0.0046 9.8i 11.0i 0.0073i 1.50 503 57.4 56.1 1.02
T30-0.94% 232 47.0 0.0053 9.5i 11.3i 0.0036i 0.94 494 47.9 43.5 1.10
T30-1.50% 232 47.0 0.0053 9.5i 11.3i 0.0036i 1.50 503 58.4 60.3 0.97
Yang et al. NR-1,2 197 46.8 0.0035g 10.7j 10.7j 0.0070j 0.00 N=A 59.3 71.0 0.84
(2010)a NR12-1,2 191 46.4 0.0035g 10.7j 10.7j 0.0070j 0.60 510j 85.1 85.1 1.00l
R13-1,2 192 46.7 0.0035g 10.7j 10.7j 0.0070j 0.09 510j 97.9 102.0 0.96
R13C-1,2 192 46.7 0.0035g 10.7j 10.7j 0.0070j 0.90 510j 97.9 92.2 1.06
R14-1,2 196 45.5 0.0035g 10.7j 10.7j 0.0070j 1.20 510j 110.6 116.7 0.95
R22-1,2 191 46.4 0.0035g 10.7j 10.7j 0.0070j 1.31 510j 105.5 106.4 0.99
R23-2 196 45.5 0.0035g 10.7j 10.7j 0.0070j 1.96 510j 128.2 131.6 0.97
Meade and S1-1 194 52.5 0.0035g 7.1j 7.1j 0.0045j 0.41 476 97.5 102.9 0.95
Graybeal S1-2 194 52.5 0.0035g 7.1j 7.1j 0.0045j 0.55 476 103.6 103.5 1.00l
(2010)d S1-3 194 52.5 0.0035g 7.1j 7.1j 0.0045j 0.64 463 113.7 121.0 0.94
S1-4 194 52.5 0.0035g 7.1j 7.1j 0.004j 0.83 476 122.4 123.5 0.99
S1-5 194 52.5 0.0035g 7.1j 7.1j 0.0045j 1.00 448 127.5 135.5 0.94
S2-0 203 53.3 0.0035g 10.7j 10.7j 0.0045j 0.00 N=A 101.5 90.0 1.13
S2-1 203 53.3 0.0035g 10.7j 10.7j 0.0045j 0.41 476 126.0 133.8 0.94l
S2-2 203 53.3 0.0035g 10.7j 10.7j 0.0045j 0.55 463 135.4 120.7 1.12
S2-3 203 53.3 0.0035g 10.7j 10.7j 0.0045j 0.64 476 141.2 163.4 0.86
S2-4 203 53.3 0.0035g 10.7j 10.7j 0.0045j 0.83 476 153.0 132.3 1.16
S2-5 203 53.3 0.0035g 10.7j 10.7j 0.0045j 1.00 448 156.7 171.2 0.91
Hasgul et al. B1-F 157 49.1f 0.0035g 6.5j 6.5j 0.0043j 0.90 453 52.6 52.6 1.00l
(2018)e B2-F 167 50.1f 0.0035g 6.5j 6.5j 0.0043j 1.90 463 80.0 89.8 0.89
B3-F 157 49.1f 0.0035g 6.5j 6.5j 0.0043j 2.80 456 100.9 111.9 0.90
B4-F 166 50.0f 0.0035g 6.5j 6.5j 0.0043j 4.30 465 98.9 134.4 0.74
Graybeal (2006b)a 80F 200 52.4 0.0043 10.0k 11.2k 0.0047k —m 1862 3,492 4,802 0.73
Graybeal (2009) a
P2-70F 221 56.3 0.0035 g
10.0 k
11.2 k
0.0047 k
— m
1862 3,616 4,248 0.85
Note: 25.4 mm ¼ 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.

© ASCE 04022013-15 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


average predicted-over-experimental moment ratio, M n =M exp , of localization crack to excessive strains and eventual rupture.
0.96 with a COV of 19.6% for all beams, providing another con- For this reason, the postlocalization flexural capacity is not rec-
firmation of the capability of the proposed methodology to predict ommended for use in design. More research is needed to quan-
the flexural capacity of prestressed and non-prestressed UHPC tify the relationship between the rupture strain of the flexural
beams. The proposed closed-form model inherently depends on the reinforcement after localization, the opening of the localized
input mechanical properties; more rigorous determination of these crack, and the development length of the reinforcement in
properties will result in a more accurate approximation of the flexu- the vicinity of the localized crack.
ral capacity. • The localization strain of UHPC attained during the preten-
sioned girder flexural test was approximately 24% greater than
the localization strain obtained from direct tension testing, in-
Summary and Conclusions dicating that the primary tensile reinforcement may enhance
the tensile strain capacity of the UHPC. More research is needed
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

Aspects of the flexural behavior of UHPC beams are distinct to refine the understanding of this behavior. Moreover, the com-
from those commonly associated with conventional concrete pression strain limit attained during the test was approximately
beams. This paper presents a flexural design framework based 20% larger than the ultimate compression strain specified in the
on rational principles with no reliance on empirical formulations. simulations, suggesting that the UHPC in a structural element
The framework is supported by an experimental investigation of the can undergo compression strains exceeding the strain at peak
flexural behavior of a pretensioned bulb-tee bridge girder. Analyti- stress obtained from a uniaxial compression test.
cal results were compared to experimental data from prestressed • Performing a capacity-based flexural design of UHPC girders
and non-prestressed flexure beam tests found in the literature that ignores the UHPC tensile resistance may incorrectly predict
and demonstrate the capability of the proposed design method a concrete crushing failure mode at the top of the section, rather
to predict flexural capacity. Based on these investigations, the fol- than localization of cracks at the bottom, and significantly over-
lowing conclusions can be reached: estimate the curvature ductility ratio, inadvertently resulting in a
• The flexural behavior of UHPC beams can be obtained from a structure that does not demonstrate the desired ductility.
sectional analysis based on a linear strain distribution in a do- • The results of the proposed design methodology were validated
main bounded by the UHPC compression strain limit, the for a pretensioned girder tested by the authors and the results are
UHPC localization strain limit, and the rupture strain of the in good agreement with the experimental data. The predicted-
reinforcing steel. The stresses on strained cross sections can be over-experimental flexural moment ratios were 0.97, 1.09, 0.99,
obtained from appropriate uniaxial stress-strain models for and 0.92 at first crack, 80% of yield, yield, and crack localiza-
UHPC, which are different from those traditionally used for tion, respectively. Additionally, the model nominal moment pre-
conventional concrete. dictions for 40 prestressed and non-prestressed beams found in
• For flexural design, the UHPC uniaxial compression stress- the literature are in good agreement with the experimental data.
strain behavior can be idealized by an elastic-perfectly plastic The average predicted-over-experimental ratio was 0.96 with a
stress-strain model. The model requires three material proper- COV of 19.6% for all beams.
ties: the modulus of elasticity, the compressive strength, and
the strain at compression strength (compression strain limit).
These properties can be obtained from compression tests per- Appendix. Supplemental Background Information
formed on cylindrical specimens according to ASTM C1856
(ASTM 2017). Flexural Design of Concrete Members
• For flexural design, the tensile stress-strain response of UHPC Flexural design of concrete members in the United States is
can be idealized by an elastic-perfectly plastic or bilinear stress- founded on mechanical models and sectional design methods speci-
strain relationship. The model requires three material property fied in AASHTO LRFD BDS (AASHTO 2020) and ACI 318-19
parameters: the cracking stress, the localization stress, and the (ACI 2019). In these documents, the analysis of the beam is based
localization strain. These properties can be obtained by execut- on the concepts of equilibrium and strain compatibility utilizing
ing a direct tension test on UHPC prismatic specimens. mechanical models to determine the stresses on strained cross sec-
• The maximum moment capacity of UHPC beams can occur at tions. In compression, the stress-strain behavior of concrete at ul-
crack localization when the strain in the extreme tensile layer is timate limit state may be simplified by a calibrated rectangular
equal to the UHPC localization strain, or at compression crush- stress block. At ultimate loading, discrete longitudinal steel rein-
ing prior to localization when the strain in the extreme compres- forcements are added and proportioned to carry all tensile stresses
sion layer is equal to the UHPC ultimate compression strain. In while ignoring the minimal tensile resistance offered by the con-
each case, the opposing criteria must not be exceeded when the crete. When subjected to flexural loads, the tensile reinforcements
primary criteria is reached. span the concrete cracks and enable flexural tensile strains far
• For members failing at crack localization, the nominal moment– greater than that at first concrete cracking. The ultimate capacity
sectional curvature diagram can be idealized by four key points of the beam is commonly limited by the attainment of a flexural
corresponding to the initiation of flexural cracks, yielding of the compression failure, defined when the strain in the concrete in com-
bottom layer of reinforcement, crack localization, and concrete pression reaches the assumed compression strain limit of 0.003.
crushing. For members failing in compression prior to localiza- Reinforced concrete flexural elements are described as (1) com-
tion, the moment-curvature diagram is simplified into the two or pression-controlled when the net tensile strain in the extreme ten-
three branches limited by concrete crushing, which can occur sion steel at ultimate is less than a specified value, generally
before or after the tensile yielding of the reinforcement. associated with the yield strain; (2) tension-controlled when the
• After localization in flexure, beams made with UHPC are ex- steel strain at failure is greater than a specified tensile strain limit
pected to develop one dominant crack. This behavior will cause value, e.g., 0.005 for non-prestressed reinforcement with a speci-
the girder to hinge at the cross section of the localized crack, fied minimum yielding stress not exceeding 517 MPa (75 ksi)
subjecting the tensile steel reinforcement crossing the (AASHTO 2020); and (3) within the transition region when the

© ASCE 04022013-16 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


net tensile strain in the steel at failure is between the specified case, the design tensile strength is taken as a reduced value of the
thresholds of the compression- and tension-controlled sections. elastic limit (cracking stress). The uniaxial tensile stresses and
The steel tensile threshold value for the tension-controlled sections strains of UHPC are obtained either from a direct tension test or
is expected to be sufficiently large to allow for excessive deforma- from an inverse analysis of the load-deflection results of flexural
tion and cracking, providing for a ductile behavior and ample prism tests. The inverse analysis method assumes a stress distribu-
warning of impending failure. In contrast, given the small tensile tion in the section at peak force and uses simplified procedures to
strains in compression-controlled sections, this failure mechanism obtain the strain-hardening portion of the UHPC stress-strain re-
may be brittle, with little warning of impending failure. Moreover, sponse based on the measured vertical deflection and sectional cur-
the failure of common compression members, i.e., columns, can vatures of the prism during the test. The applicability of the Swiss
lead to detrimental consequences to the stability of the entire design recommendation document is limited to UHPC materials
structure. To address this issue, a reduced moment resistance factor exhibiting a uniaxial stress-strain response showing a continuous
is specified for compression-controlled sections, i.e., ϕf ¼ 0.75, increase in stress after cracking until the peak stress is reached,
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

compared to the resistance factor specified for tension-controlled i.e., strain-hardening behavior.
sections, i.e., ϕf ¼ 0.90, for non-prestressed and prestressed In the Canadian design recommendations, the design capacity of
components designed according to ACI 318-19 (ACI 2019), and flexural elements composed of strain-hardening fiber-reinforced
ϕf ¼ 0.90 and 1.00 for non-prestressed components and pre- concrete (SH-FRC), such as UHPC, is calculated by determining the
stressed components designed according to AASHTO LRFD BDS sectional forces corresponding to a linear strain diagram bounded by
(AASHTO 2020), respectively. 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
International Recommendations for Flexural Design value in the extreme compression layer is less than the limiting strain
In the French design standard, the resisting forces of the sections of at crushing. When the extreme compression fiber of the member
a member subjected to flexure are calculated from a linear strain equals the ultimate compression strain (i.e., crushing limit), the com-
diagram included in a domain bounded by a compression strain pression model can be taken as the traditional rectangular stress
limit, corresponding to UHPC crushing, and a reinforcing steel block, which is compatible with a triangular stress distribution for
limit, corresponding to the rupture of reinforcing steel (AFNOR materials with compression strength values exceeding 120 MPa
2016a). For members with no steel reinforcing bars or prestressing (17.4 ksi). The tensile model at the ultimate limit state is a rectan-
strands, tensile strains are limited to the strain at peak tensile gular distribution extending over cross-sectional layers where the
strength of UHPC. The cross-sectional forces are obtained by em- tensile strains are less than or equal to the limiting tensile strain
ploying compression and tension mechanical stress-strain models to (i.e., strain at peak tensile resistance). The compression parameters
calculate the stresses corresponding to the assumed strains. The are obtained from uniaxial compression tests, while the uniaxial ten-
compression models include a detailed model based on parabolic sile parameters (i.e., peak stress and strain) are obtained either from a
representation of the stress-strain results obtained from compres- direct tension test or from an inverse analysis of the load-deflection
sion tests and a simplified elastic-perfectly plastic model with re- results of flexural prism tests. The inverse analysis technique is
duced design stress and strain limits. The tension models include based on empirical approximations of the stresses at specific vertical
an elastic-perfectly plastic or bilinear stress-strain representation, deflections and crack openings recorded during the testing of a
limited by the strain at crack localization (hardening limit), for flexural prism of specified dimensions. The Canadian design docu-
materials exhibiting a constant or increasing stress after cracking, ment specifies a flexural resistance method based on sectional cur-
respectively. The material compression parameters are obtained vature ductility. The maximum flexural resistance M1 is taken as the
directly from uniaxial compression tests, while the tensile param- maximum of the flexural moment (1) at yielding of the reinforce-
eters are obtained indirectly from a prism flexure test using an ment in tension, (2) when the strain in the extreme tension fiber is
inverse analysis technique. The inverse analysis back-calculates equal to the limiting tensile strain, or (3) at concrete crushing. The
an approximation of the uniaxial tensile stress-strain response curvature ductility ratio is defined as the ratio of the sectional cur-
from the moment-deflection results of the prism test and includes vature at M 1 to the sectional curvature at the yielding of the tensile
specific procedures to infer the sectional curvatures of the beam flexural reinforcement. If the ductility ratio is greater than 2.0, the
corresponding to the measured deflections at each load step flexural resistance of the member is taken equal to the moment at
(AFNOR 2016b). yielding of reinforcement. In cases where the curvature ductility is
In the Swiss design recommendations, the ultimate bending re- less than 2.0, the flexural resistance of the member must be taken as
sistance of a reinforced UHPC cross section is calculated by resolv- the maximum of 0.50 × M 1 or the flexural resistance calculated
ing the stresses corresponding to a linear strain diagram included in without considering the contribution of fibers. Fig. 1 presents a com-
a domain bounded by the strains when the peak compression parison of the cross-sectional stress conditions of UHPC members
(crushing) and tensile (localization) strengths are reached (SIA computed following the recommendations of the French, Swiss, and
2016). The compression stress-strain model is linear elastic to fail- Canadian documents when the strain at the extreme tension layer is
ure, with the failure stress defined by the design compression equal to the design tensile strain limit εt;lim and with elastic stresses
strength obtained from compression testing of cylinders or cubes. in compression.
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 0.9ðh − XÞ from Data Availability Statement
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 Some or all data, models, or code that support the findings of this
to describe a UHPC member with steel reinforcing bars or strands study are available from the corresponding author upon reasonable
as the primary flexural tensile reinforcement, the tensile strain limit request. Some or all data, models, or code generated or used during
(strain at peak tensile strength) may be increased by a factor of 2.0 the study are proprietary or confidential in nature and may only be
due to simultaneous action of the fibers and reinforcing steel; in this provided with restrictions.

© ASCE 04022013-17 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


Acknowledgments M sl = baseline flexural moment for ductility considerations,
calculated when the bottom layer of steel is equal to 80%
The research presented in this paper was funded by FHWA. The the yield stress of steel reinforcement;
publication of this paper does not necessarily indicate approval M y = flexural moment at yielding of the extreme layer of
or endorsement of the findings, opinions, conclusions, or recom- reinforcement;
mendations either inferred or specifically expressed herein by M1 = maximum flexural moment as defined by Section
FHWA or the United States Government. This research could CA8.1.8.4.3.1 of CSA S6:19 documents;
not have been completed were it not for the dedicated support X = distance from extreme compression fiber to the neutral
of the technical professionals associated with the FHWA Structural axis of the section;
Concrete Research Program. α = reduction factor on the compressive strength reflecting
the linearity limit of the material compressive stress-
strain response;
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

Notation γ = reduction factor applied on the tensile stress parameters


(ft;cr and ft;loc ) to account for variability in the material
The following symbols are used in this paper: testing results;
d = distance from top of the section to the centroid of ΔX = distance from extreme compression fiber to the neutral
reinforcing steel; axis of the section of the intermediate strain profile in
Ec = modulus of elasticity of UHPC; Fig. 13;
Ēc = average modulus of elasticity of UHPC; ΔεB = change in strain at midspan at extreme bottom fiber;
Ep = modulus of elasticity of prestressing steel strands; Δεcp1 = change in strain in the concrete at midspan at the level of
Es = modulus of elasticity of conventional steel; the first layer of strands;
fc = stress in extreme compression layer; Δεcp2 = change in strain in the concrete at midspan at the level of
fc0 = compressive strength of UHPC; the second layer of strands;
f̄c0 = average compressive strength of UHPC; Δεcp3 = change in strain in the concrete at midspan at the level of
fcrd;F = design cracking strength of UHPC as defined by the third layer of strands;
Fig. 3.206 of NF P18-710; Δεi = change in strain in UHPC layer i;
fps = stress in prestressing strand layer; ΔεT = change in strain at midspan at extreme top fiber;
f pu = ultimate tensile strength of prestressing strands; Δψ = change in sectional curvature taken as Δεt =ΔX;
f̄ pu = average ultimate tensile strength of prestressing εB = strain at midspan at extreme bottom fiber;
strands; εB;ini = initial strain at midspan at extreme bottom fiber,
f py = yield strength of prestressing strands; corresponding to σB;ini ;
f̄ py = average yield strength of prestressing strands; εc = strain in extreme compression layer;
fs = stress in reinforcing steel layer; εcp = elastic compression strain limit taken as αf c0 =E;
fs;d = design stress of reinforcing steel as defined by the εcp1 = strain in the concrete at midspan at the level of the first
appropriate international design document; (top) layer of strands;
f sy = yield stress of conventional steel; εcp2 = strain in the concrete at midspan at the level of the
f t;r = effective cracking stress of UHPC; second layer of strands;
f̄t;cr = average effective cracking stress of UHPC; εcp3 = strain in the concrete at midspan at the level of the third
f td;C = design tensile strength of UHPC as defined by Fig. (bottom) layer of strands;
CA8.1.8a of CSA S6:19; εcp1;ini = initial strain in the UHPC at the level of the first (top)
ftd;F = design tensile strength of UHPC as defined by Fig. 3.206 layer of prestressing strands;
of NF P18-710; εcp2;ini = initial strain in the UHPC at the level of the second layer
f td;S = design tensile strength of UHPC as defined by Section of prestressing strands;
2.4.2.3 of SIA 2052; εcp3;ini = initial strain in the UHPC at the level of the third
f t;lim = tensile strain limit under service loads; (bottom) layer of prestressing strands;
ft;loc = localization stress of UHPC; εcu = strain at ultimate compressive strength of UHPC;
f̄t;loc = average localization stress of UHPC; ε̄cu = average strain at ultimate compressive strength of
h = overall thickness or height of a member; UHPC;
M = applied bending moment on UHPC girder; εi = strain in UHPC layer i;
Mc = flexural moment at crushing of UHPC at extreme εi;ini = initial strain in UHPC layer i;
compression layer; εpe = strain in the effective prestress;
M cr = flexural moment at the first flexural crack; εps = strain in the prestressing strand layer;
M exp = total experimental moment corresponding to applied εpu = rupture tensile rupture strain of prestressing strands;
moment, M, plus the self-weight moment, M ini ; ε̄pu = average tensile rupture strain of prestressing
Mini = bending moment induced by the self-weight of the strands;
UHPC girder and loading apparatus; εpy = yield strain of prestressing strands;
M L = flexural moment at crack localization, when the extreme εp1 = strain in the first (top) layer of steel strands;
tensile layer, εt , is equal to the localization strain of εp2 = strain in the second layer of steel strands;
UHPC, εt;loc ; εp3 = strain in the third (bottom) layer of steel strands;
M n = nominal flexural moment; εp1;ini = initial strain in the first (first) layer of prestressing strands
Mr = factored flexural moment; (at the star M ini );

© ASCE 04022013-18 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


εp2;ini = initial strain in the second layer of prestressing strands ASTM. 2014. Standard test method for static modulus of elasticity and
(strain at M ini ); poisson’s ratio of concrete in compression. ASTM C469/C469M. West
εp3;ini = initial strain in the third (bottom) layer of prestressing Conshohocken, PA: ASTM.
ASTM. 2017. Standard practice for fabricating and testing specimens
strands (strain at M ini );
of ultra-high performance concrete. ASTM C1856/C1856M. West
εs = strain in reinforcing steel layer; Conshohocken, PA: ASTM.
εsu = rupture strain of conventional steel; ASTM. 2018a. Standard specification for low-relaxation, seven-wire steel
εsy = yield strain of conventional steel; strand for prestressed concrete. ASTM A416/A416M. West Consho-
εT = strain at midspan at extreme top fiber; hocken, PA: ASTM.
εT;ini = initial strain at midspan at extreme top fiber, ASTM. 2018b. Standard test method for compressive strength of cylindri-
corresponding to σT;ini ; cal concrete specimens. ASTM C39/C39M. West Conshohocken, PA:
εt = strain in extreme tensile layer; ASTM.
ASTM. 2020. Standard test method and definitions for mechanical testing
εt;cr = strain at effective cracking stress of UHPC;
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

of steel products. ASTM A370. West Conshohocken, PA: ASTM.


εt;lim = design tensile strain limit of UHPC as defined by the Baby, F., P. Marchand, and F. Toutlemonde. 2014. “Shear behavior of ultra-
appropriate international design document; high performance fiber-reinforced concrete beams. I: Experimental in-
εt;loc = average localization strain of UHPC; vestigation.” J. Struct. Eng. 140 (5): 04013111. https://doi.org/10.1061
ε̄t;loc = average localization strain of UHPC; /(ASCE)ST.1943-541X.0000907.
μ = curvature ductility ratio defined as ψL =ψy ; Bencardino, F., L. Rizzuti, G. Spadea, and R. N. Swamy. 2010. “Experi-
μexp = experimental value of curvature ductility ratio; mental evaluation of fiber reinforced concrete fracture properties.” Com-
posites, Part B 41 (1): 17–24. https://doi.org/10.1016/j.compositesb
μmin = specified minimum curvature ductility ratio;
.2009.09.002.
μnt;sim = simulated value of curvature ductility ratio when the Bolander, J. E., S. Choi, and S. R. Duddukuri. 2008. “Fracture of fiber-
tensile resistance of UHPC is ignored; reinforced cement composites: Effects of fiber dispersion.” Int. J. Fract.
μsim = simulated value of curvature ductility ratio; 154 (1–2): 73–86. https://doi.org/10.1007/s10704-008-9269-4.
ρ = reinforcement ratio of conventional steel; Chen, L., and B. A. Graybeal. 2012. “Modeling structural performance
σB;ini = initial stress at midspan at extreme bottom fiber; of ultrahigh performance concrete I-girders.” J. Bridge Eng. 17 (5):
σT;ini = initial stress at midspan at extreme top fiber; 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
ϕf = resistance factor for flexural moment;
rebar-reinforced ultra-high-performance concrete beams.” Mag. Concr.
ψ = sectional curvature defined as the ration of the Res. 70 (19): 997–1015. https://doi.org/10.1680/jmacr.17.00283.
compressive strain in the section over the depth of CSA (Canadian Standard Association). 2019. Canadian highway bridge
the neutral axis; design code. S6:19. Torronto, ON: CSA.
ψc = sectional curvature at crushing of UHPC at extreme Cunha, V. M. C. F., J. A. O. Barros, and J. M. Sena-Cruz. 2011. “An in-
compression layer; tegrated approach for modelling the tensile behaviour of steel fibre re-
ψcr = sectional curvature at the first flexural crack; inforced self-compacting concrete.” Cem. Concr. Res. 41 (1): 64–76.
ψL = sectional curvature at crack localization, when the https://doi.org/10.1016/j.cemconres.2010.09.007.
extreme tensile layer, εt , is equal to the localization strain De Larrard, F., and T. Sedran. 2002. “Mixture-proportioning of high-
of UHPC, εt;loc ; performance concrete.” Cem. Concr. Res. 32 (11): 1699–1704. https://
doi.org/10.1016/S0008-8846(02)00861-X.
ψsl = baseline sectional curvature for ductility considerations,
Doyon-Barbant, J., and J. P. Charron. 2018. “Impact of fibre orientation on
calculated when the bottom layer of steel is equal to 80%
tensile, bending and shear behaviors of a steel fibre reinforced con-
the yield stress of steel reinforcement;
crete.” Mater. Struct. 51 (6): 157. https://doi.org/10.1617/s11527-018
ψy = sectional curvature at yielding of the extreme layer of -1282-0.
reinforcement; and El-Helou, R., and B. Graybeal. 2019. “The ultra girder: A design concept
ω̄c = average density of UHPC. 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.
References El-Helou, R. G. 2016. “Multiscale computational framework for analysis
and design of ultra-high performance concrete structural components
Aaleti, S., B. Petersen, and S. Sritharan. 2013. Design guide for and systems.” Ph.D. dissertation, Dept. of Civil and Environmental En-
precast UHPC Waffle Deck panel system, including connections. gineering, Virginia Polytechnic Institute and State Univ.
Rep. No. FHWA-HIF-13-032. Washington, DC: Federal Highway El-Helou, R. G., Z. B. Haber, and B. A. Graybeal. 2022. “Mechanical
Administration. behavior and design properties of ultra-high performance concrete.”
AASHTO. 2020. LRFD bridge design specifications. 9th ed. Washington,
ACI Mater. J. 119 (1). https://doi.org/10.14359/51734194.
DC: AASHTO.
El-Helou, R. G., I. Koutromanos, C. D. Moen, and M. Moharrami. 2020.
AASHTO. 2022. Standard method of test for uniaxial tensile response of
“Triaxial constitutive law for ultra-high-performance concrete and
ultra-high performance concrete. AASHTO T 397 Washington, DC:
AASHTO. other fiber-reinforced cementitious materials.” J. Eng. Mech. 146 (7):
ACI (American Concrete Institute). 2019. Building code requirements for 04020062. https://doi.org/10.1061/(ASCE)EM.1943-7889.0001777.
structural concrete and commentary. ACI 318-19. Farmington Hills, Gowripalan, N., and I. R. Gilbert. 2000. Design guidelines for ductal pre-
MI: ACI. stressed concrete beams. Sydney: The University of New South Wales,
AFNOR (Association Française de Normalisation). 2016a. National addi- VSL (Aust) Pty Ltd.
tion to the Eurocode 2—Design of concrete structures: Specific rules Graybeal, B. A. 2006a. Material property characterization of ultra-high
for ultra-high performance fibre-reinforced concrete (UHPFRC). NF performance concrete. Rep. No. FHWA-HRT-06-103. McLean, VA:
P18-710. Paris: AFNOR. Federal Highway Administration.
AFNOR (Association Française de Normalisation). 2016b. Ultra-high Graybeal, B. A. 2006b. Structural behavior of ultra-high performance con-
performance fibre-reinforced concrete—Specifications, performance, crete prestressed I-girders. Rep. No. FHWA-HRT-06-115. McLean,
production, and conformity. NF P18-470. Paris: AFNOR. VA: Federal Highway Administration.

© ASCE 04022013-19 J. Struct. Eng.

J. Struct. Eng., 2022, 148(4): 04022013


Graybeal, B. A. 2008. “Flexural behavior of an ultrahigh-performance con- Qiu, M., Y. Zhang, S. Qu, Y. Zhu, and X. Shao. 2020b. “Effect of reinforce-
crete I-girder.” J. Bridge Eng. 13 (6): 602–610. https://doi.org/10.1061 ment ratio, fiber orientation, and fiber chemical treatment on the direct
/(ASCE)1084-0702(2008)13:6(602). tension behavior of rebar-reinforced UHPC.” Constr. Build. Mater.
Graybeal, B. A. 2009. Structural behavior of a prototype ultra-high per- 256 (Sep): 119311. https://doi.org/10.1016/j.conbuildmat.2020.119311.
formance concrete pi-girder. Report No. FHWA-HRT-10-027. McLean, Radtke, F. K. F., A. Simone, and L. J. Sluys. 2010. “A computational model
VA: Federal Highway Administration. for failure analysis of fibre reinforced concrete with discrete treatment
Graybeal, B. A., and F. Baby. 2013. “Development of direct tension test of fibres.” Eng. Fract. Mech. 77 (4): 597–620. https://doi.org/10.1016/j
method for ultra-high- performance fiber-reinforced concrete.” ACI .engfracmech.2009.11.014.
Mater. J. 110 (2): 177–186. Russel, H., and B. A. Graybeal. 2013. Ultra-high performance concrete:
Graybeal, B. A., and B. Stone. 2012. Compression response of a rapid- A state-of-the-art report for the bridge community. Rep. No. FHWA-
strengthening ultra-high performance concrete formulation. Rep. No. HRT-13-060. McLean, VA: Federal Highway Administration.
FHWA-HRT-12-064. McLean, VA: Federal Highway Administration. Schauffert, E. A., and G. Cusatis. 2012. “Lattice discrete particle model for
Habel, K., M. Viviani, E. Denarié, and E. Brühwiler. 2006. “Development fiber-reinforced concrete. I: Theory.” J. Eng. Mech. 138 (7): 826–833.
Downloaded from ascelibrary.org by Federal Highway Administration on 01/31/22. Copyright ASCE. For personal use only; all rights reserved.

of the mechanical properties of an ultra-high performance fiber rein- https://doi.org/10.1061/(ASCE)EM.1943-7889.0000387.


forced concrete (UHPFRC).” Cem. Concr. Res. 36 (7): 1362–1370. Shao, Y., and S. L. Billington. 2019. “Utilizing full UHPC compressive
https://doi.org/10.1016/j.cemconres.2006.03.009. strength in steel reinforced UHPC beams.” In Proc., 2nd Int. Interactive
Haber, Z., I. De La Varga, B. Graybeal, B. Nakashoji, and R. El-Helou. Symp. on Ultra-High Performance Concrete. Ames, IA: Iowa State
2018. Properties and behavior of UHPC-class materials. Rep. No. Univ.
FHWA-HRT-18-036. McLean, VA: Federal Highway Administration. SIA (Swiss Society of Engineers and Architects). 2016. Recommendation:
Hasgul, U., K. Turker, T. Birol, and A. Yavas. 2018. “Flexural behavior of Ultra-high performance fibre-reinforced cement-based composites
ultra-high-performance fiber reinforced concrete beams with low and (UHPFRC) construction material, dimensioning, and application.
high reinforcement ratios.” Struct. Concr. 19 (6): 1577–1590. https:// 2052. Lausanne, Switzerland: SIA.
doi.org/10.1002/suco.201700089. Skogman, B. C., M. K. Tadros, and R. Grasmick. 1988. “Flexural strength
Huang, H., X. Gao, L. Li, and H. Wang. 2018. “Improvement effect of of prestressed concrete members.” PCI J. 33: 96–123.
steel fiber orientation control on mechanical performance of UHPC.” Smith, J., G. Cusatis, D. Pelessone, E. Landis, J. O’Daniel, and J. Baylot.
Constr. Build. Mater. 188 (Nov): 709–721. https://doi.org/10.1016/j 2014. “Discrete modeling of ultra-high-performance concrete with ap-
.conbuildmat.2018.08.146. plication to projectile penetration.” Int. J. Impact Eng. 65 (Mar): 13–32.
Lepech, M. D., and V. C. Li. 2009. “Water permeability of engineered ce- https://doi.org/10.1016/j.ijimpeng.2013.10.008.
mentitious composites.” Cem. Concr. Compos. 31 (10): 744–753. Soranakom, C., and B. Mobasher. 2008. “Correlation of tensile and flexu-
https://doi.org/10.1016/j.cemconcomp.2009.07.002. ral responses of strain softening and strain hardening cement compo-
Li, C., Z. Feng, L. Ke, R. Pan, and J. Nie. 2019. “Experimental study on sites.” Cem. Concr. Compos. 30 (6): 465–477. https://doi.org/10.1016
shear performance of cast-in-place ultra-high performance concrete /j.cemconcomp.2008.01.007.
structures.” Materials (Basel) 12 (19): 3254. https://doi.org/10.3390 Soranakom, C., and B. Mobasher. 2009. “Flexural design of fiber-
/ma12193254. reinforced concrete.” ACI Mater. J. 106 (5): 461–469.
Magureanu, C., I. Sosa, C. Negrutiu, and B. Heghes. 2012. “Mechanical Stürwald, S. 2017. “Bending behaviour of UHPC reinforced with rebars
properties and durability of ultra-high-performance concrete.” ACI and steel fibres.” In Proc., 2017 fib Symp.: High Tech Concrete: Where
Mater. J. 109 (2): 177–184. Technology and Engineering Meet, 473–481. Cham, Switzerland:
Maya Duque, L. F., I. De La Varga, and B. Graybeal. 2016. “Fiber Springer.
reinforcement influence on the tensile response of UHPFRC.” In Proc., Walsh, K. K., N. J. Hicks, E. P. Steinberg, H. H. Hussein, and A. A.
1st Int. Interactive Symp. on UHPC. Ames, IA: Iowa State Univ. Semendary. 2018. “Fiber orientation in ultra-high-performance concrete
Meade, T., and B. Graybeal. 2010. “Flexural response of lightly reinforced shear keys of adjacent-box-beam bridges.” ACI Mater. J. 115 (2): 227–
ultra-high performance concrete beams.” In Proc., 3rd fib Int. Congress. 238. https://doi.org/10.14359/51701097.
Washington, DC: Fédération International du béton. Wille, K., and A. E. Naaman. 2013. “Effect of ultra-high-performance con-
NJDOT (New Jersey DOT). 2016. Design manual for bridges and struc- crete on pullout behavior of high-strength brass-coated straight steel
tures. 6th ed. Trenton, NJ: NJDOT. fibers.” ACI Mater. J. 110 (4): 451–462.
Oesterlee, C. 2010. “Structural response of reinforced UHPFRC and RC Yang, I. H., C. Joh, and B. S. Kim. 2010. “Structural behavior of ultra high
composite members.” Ph.D. dissertation, School of Architecture, Civil performance concrete beams subjected to bending.” Eng. Struct.
and Environmental Engineering, Swiss Federal Institute of Technology 32 (11): 3478–3487. https://doi.org/10.1016/j.engstruct.2010.07.017.
Lausanne. Yao, Y., X. Wang, K. Aswani, and B. Mobasher. 2017. “Analytical proce-
PCI (Precast/Prestressed Concrete Institute). 2014. PCI bridge design dures for design of strain softening and hardening cement composites.”
manual. 3rd ed. Second Release. Chicago: PCI. Int. J. Adv. Eng. Sci. Appl. Math. 9 (3): 181–194. https://doi.org/10
Peng, X., and C. Meyer. 2000. “A continuum damage mechanics model .1007/s12572-017-0187-4.
for concrete reinforced with randomly distributed short fibers.” Yin, H., K. Shirai, and W. Teo. 2019. “Finite element modelling to predict
Comput. Struct. 78 (4): 505–515. https://doi.org/10.1016/S0045-7949 the flexural behaviour of ultra-high performance concrete members.”
(00)00045-6. Eng. Struct. 183 (Mar): 741–755. https://doi.org/10.1016/j.engstruct
Pros, A., P. Diez, and C. Molins. 2012. “Modeling steel fiber reinforced .2019.01.046.
concrete: Numerical immersed boundary approach and a phenomeno- Yoo, D. Y., N. Banthia, and Y. S. Yoon. 2017. “Experimental and numerical
logical mesomodel for concrete-fiber interaction.” Int. J. Numer. Meth- study on flexural behavior of ultra-high-performance fiber-reinforced
ods Eng. 90 (1): 65–86. https://doi.org/10.1002/nme.3312. concrete beams with low reinforcement ratios.” Can. J. Civ. Eng.
Qiu, M., X. Shao, K. Wille, B. Yan, and J. Wu. 2020a. “Experimental in- 44 (1): 18–28. https://doi.org/10.1139/cjce-2015-0384.
vestigation on flexural behavior of reinforced ultra high performance Yoo, D. Y., and Y. S. Yoon. 2015. “Structural performance of ultra-high-
concrete low-profile T-beams.” Int. J. Concr. Struct. Mater. 14 (1): performance concrete beams with different steel fibers.” Eng. Struct.
5. https://doi.org/10.1186/s40069-019-0380-x. 102 (Nov): 409–423. https://doi.org/10.1016/j.engstruct.2015.08.029.

© ASCE 04022013-20 J. Struct. Eng.

View publication stats J. Struct. Eng., 2022, 148(4): 04022013

You might also like