You are on page 1of 10

Engineering Structures 164 (2018) 109–118

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Design of anchor reinforcement for seismic tension loads T


a b c,⁎
Derek Petersen , Zhibin Lin , Jian Zhao
a
Osmose Railroad Services Inc., Madison, WI 57716, USA
b
Department of Civil and Environmental Engineering, North Dakota State University, Fargo, ND 58108, USA
c
Department of Civil and Environmental Engineering, University of Wisconsin, Milwaukee, WI 53201, USA

A R T I C L E I N F O A B S T R A C T

Keywords: This paper presents a study of cast-in anchors in reinforced concrete subjected to both monotonic and cyclic
Cast-in anchors tension. Laboratory tests using 25 mm [1 in.] diameter anchor bolts showed that the code-conforming anchor
Headed studs reinforcement may not ensure anchor steel failure though it effectively prevented concrete breakout. Based on
Anchor connections the observations and other tests in the literature, recommendations for the design of anchor tension reinforce-
Anchor reinforcement
ment were proposed, which consists of (1) load-carrying reinforcement in the direction of the anchors; (2) crack-
Fastening to concrete
controlling reinforcement in all directions that have a limited edge distance; and (3) local confining re-
Composite construction
And seismic design inforcement near the anchor head if side-face blowout may control the failure. The proposed anchor re-
inforcement was proven effective using additional laboratory tests. Further studies are needed to verify the
reinforcement design for anchor connections.

1. Introduction before the reinforcement takes effect. With the cracks, the tensile ca-
pacity of the anchors is provided by the reinforcement in the same di-
Anchor connections are a critical component of load transfer be- rection the anchors, and the tensile capacity of the reinforcement can be
tween steel and concrete members affecting structural performance affected by available development lengths [10,11].
during earthquake events [1,2]. Cast-in anchors may experience steel Anchor reinforcement in this study is that specifically designed and
fracture and/or concrete breakout failure when subjected to a tension detailed such that the reinforced anchors achieve their full design ca-
force [3–5]. The failure modes are mainly dependent upon the em- pacities. U-shaped hairpins are specified in the current building codes
bedment depth (hef) when the anchor bolt is placed in plain concrete. and design guides. In addition, reinforcing bars placed near anchor
Breakout failure is brittle and thus not a preferred failure mode for the bolts have been expected to create loading transferring paths through a
anchors in seismic zones [6]. Engineers are recommended to consider splice mechanism (Fig. 1c, e, and f), and explained using strut-and-tie
steel reinforcement by building codes [7] and design guides [8,9] to models (STMs). Few experimental tests have been available to verify
improve the behavior of anchors likely controlled by concrete breakout. the effectiveness of these anchor reinforcement, especially under si-
It has been well documented that the potential failure region near mulated seismic loading.
headed stud anchors be confined using steel reinforcement [10]. The
existing reinforcement for concrete anchors in the technical literature is 2. Literature review
summarized in Fig. 1. Two types of reinforcement are allowed in the
building codes [7] and design guides [8]: specifically designed re- Very limited studies in the literature have focused on the behavior
inforcement (known as anchor reinforcement) for cast-in anchors and of anchor reinforcement and reinforced anchors. However, the litera-
the existing reinforcement near the anchors in a reinforced concrete ture review below includes a variety of related studies because concrete
member (known as supplementary reinforcement). With a goal of re- anchors are widely used in many applications in construction.
straining potential concrete breakout cracks, supplementary reinforce- Specifically, conceptual designs in the literature and the studies, in
ment is not designed to transfer the full tensile capacity of anchors into which headed anchors were embedded in reinforced concrete members,
concrete. The effect of supplementary reinforcement has been well in which some reinforcing bars may have improved the anchor beha-
studied [11,12], and the corresponding analysis methods have been vior/capacities. Observations from these tests are used to complement
proposed. For example, the existing design methods assume that the those made in this study, based which, an improved design method is
concrete breakout, shown in shaded areas in Fig. 1a, c, and d, occurs proposed.


Corresponding author.
E-mail address: jzhao@uwm.edu (J. Zhao).

https://doi.org/10.1016/j.engstruct.2018.02.015
Received 11 October 2016; Received in revised form 23 December 2017; Accepted 1 February 2018
0141-0296/ © 2018 Elsevier Ltd. All rights reserved.
D. Petersen et al. Engineering Structures 164 (2018) 109–118

foundation to the column reinforcement, as illustrated in Fig. 1e. The


Nomenclature
same model was also used to determine the lateral bursting force that
leads to side face blowout failure. In the tests of column base connec-
ca1 front edge distance of anchor
tions in foundation beams, Baba et al. [17] treated closed stirrups near
da anchor diameter
the headed anchors as anchor reinforcement, and applied STMs in ex-
db reinforcement diameter
plaining the load transfer. More importantly, a separate strut-and-tie
ds stirrup diameter
model was used to visualize the load transfer from the bottom of the
fc measured concrete compressive strength
closed stirrups to the rest of the structure: column base plate in com-
fc′ specified concrete compressive strength (27.5 MP
[4000 psi] in the study) pression, as illustrated in Fig. 1f.
fy specified yield strength of steel (414 MPa [60 ksi] in the
study) 2.2. Existing experimental studies
hef embedment depth of anchor bolts
There are limited tests available in the literature focusing on the

Fig. 1. Schematics of existing anchor tension reinforcement.

2.1. Existing design recommendations verification of the above recommended anchor reinforcement. Focusing
on the concrete breakout capacity of large diameter anchors, the tests
The recommendations on the design of anchor tension reinforce- by Lee et al. [18] included two test groups with four or eight No. 8 U-
ment are summarized in Fig. 1. Cannon et al. [13] suggested simple shaped hairpins, located within 0.2hef and the other 0.35hef from the
hairpins for preventing potential breakout cones as shown in Fig. 1a. To test anchors. The No. 8 hairpins were assumed fully developed below
achieve direct force transfer (similar to rebar splicing), hairpins are the anchor head through a length of 14db with hooked ends. Note that
required to be placed symmetrically within 1/3hef , measured from the these specimens were referred as “reinforced specimens,” which leads
edge of the anchor head, where hef is the embedded depth. A similar to the term “reinforced anchors” in this paper. The hairpins can be
recommendation was adopted by CEB (1997) and ACI (2008) with a treated as supplementary reinforcement because they were not pro-
larger effective range (0.5hef ), measured from the center of anchors, as portioned to carry the anchor steel capacity in tension. The tests in-
shown in Fig. 1b. The European committee for standardization (2009) dicate that hairpins close to the test anchor are more effective than
further increased this effective range to 0.75hef . Critical to this study, those located farther. The authors also indicated that the reinforcement
stirrups in beams have been allowed by fib (2008) to transfer tensile should be proportioned to carry 100% of the applied load, and the
forces from anchors installed at the bottom of the beam to the com- design strength for the reinforcement should consider bond failure of
pression zone of the beam, as illustrated in Fig. 1c. reinforcement in the breakout cone. Berger [12] further studied the
Cannon et al. [13] also recommended spiral reinforcement for each effect of supplementary reinforcement, and proposed a new capacity
anchor in tension to prevent “lateral bursting,” now known as side face model, which considers contribution from both concrete and steel re-
blowout failure. In an application of U-shaped hairpins to the connec- inforcement. The tensile capacity of the reinforcement was calculated
tions between steel columns and their concrete foundations, shown in based on its bond capacity. Sharma et al. [11,19] extended the in-
Fig. 1d, Shipp and Haninger [14] recommended closed ties at a small vestigation to groups of four anchors. Instead of hairpins, closed stir-
spacing (e.g. 150 mm [6 in.]) to encase the anchor bolts. This critical rups were used as supplementary reinforcement in this latest study.
measure may have been recommended after the study by Hasselwander There are tests in the literature with headed studs placed in re-
et al. [15], in which many specimens, without confining reinforcement, inforced concrete members. The existing reinforcement in the members
failed in concrete side-face blowout. was not specifically designed as anchor reinforcement; however, ob-
Strut-and-tie models (STMs) have been used in the analysis and servations from these tests provide relevant and valuable information to
design of reinforcement for concrete anchors. For example, STMs are the subject of this study. For example, closed stirrups were provided in
used to demonstrate the effectiveness of hairpins, and the needs for one of the two tension tests of headed studs for use in composite con-
crack-controlling reinforcement [4,5,7,8]. Widianto and Patel [16] ex- struction by Saari et al. [20]. The No. 3 stirrups at a spacing of 89 mm
plained the load transfer from headed anchors embedded in a column [3.5 in.] and longitudinal bars at all corners formed a reinforcing cage

110
D. Petersen et al. Engineering Structures 164 (2018) 109–118

around the studs. Each stud thus had two No. 3 stirrups (four legs as breakout cone with a radius of two times the embedment depth (2hef)
anchor reinforcement), which were able to resist a tension force higher from each anchor plus an additional 100 mm [4 in.] on both sides. The
than the stud capacity in tension. Concrete failure occurred to the un- block height was 432 mm [17 in.] similar to all other anchor tests in this
reinforced anchor while steel fracture occurred to the reinforced anchor study, which was limited by the position of the vertical actuator [6].
with capacity increase about 100 percent. Similar observation was also Five types of reinforcement patterns were tested first as shown in
made by Baran et al. [21] in their tests, in which reinforcing cages, Fig. 2, and listed in Table 1. From the measured strength of the
consisting of No. 4 stirrups at 150 mm [6 in.] and corner bars at 75 mm threaded rods shown later, the ultimate tensile capacity of the anchor
[3 in.], were provided to the studs. Compared with unreinforced spe- bolts was 358 kN [80.4 kips]. The anchor reinforcement using Grade 60
cimens, reinforced studs achieved higher capacities though their failure steel was proportioned for this load assuming a nominal yield strength
was not controlled by stud fracture. of 414 MPa [60 ksi]. The required anchor reinforcement was found to
In addition, Shahrooz et al. [22] conducted six 1/3-scale tests of be 865 mm2 [1.34 in.2], which was provided using four No. 6 bars
shear connections between concrete walls and steel outrigger beams for (Types B and D), six No. 4 and two No. 3 bars (Types C and E), or eight
use in hybrid structures. The 4-stud and 6-stud connections with 13-mm No. 4 bars (Type F) as shown in Fig. 2.
[0.5 in.] diameter headed studs were installed in concrete representing The anchor reinforcement is assumed to develop strains beyond the
the boundary elements of a concrete wall. Specimen No. 4 was used to proportional limit; therefore, the bars need to be fully developed at
simulate the boundary elements at lower levels, in which D2.9 stirrups both sides of the assumed breakout crack. All stirrups in Group 1 spe-
were provided with a spacing of 38 mm [1.5 in.] near the connection. cimens were 279 mm [11 in.] high measured from the outside edges.
The shear tab in Specimen No. 4 yielded while concrete breakout was Concrete cover of 38 mm [1.5 in.] was used, thus stirrups were em-
observed in all other fours tests. The benefit of closely spaced stirrups bedded about 102 mm [4 in.] above the anchor head. This short de-
was also observed in later tests of two ¼-scale girder-wall connections velopment length was deemed sufficient to develop the vertical legs of
in concrete experiencing damage resulted from simulated seismic ac- the closed stirrups through the interaction with the corner bars [24].
tions [23]. Similarly, the development length below the anchor head was 152 mm
This study focuses on specially designed and detailed reinforcement [6 in.].
for anchors to achieve their full capacities corresponding to steel frac-
ture in tension. Tests of anchors with code-complying anchor re- 3.2. Materials
inforcement conducted to evaluate the easing design methods. An im-
proved reinforcement design, based on observations in this study and The measured stress-strain relationship for a coupon made from
those in the literature, was proposed and experimentally verified to
ASTM A193 Grade B7 rods is shown in Fig. 3. The sudden slope change
ensure anchor steel failure in addition to restraining concrete breakout near 655 MPa [95 ksi] may have been due to a slip of the coupon out of
cracks.
the grips of the loading frame. No additional tests were conducted be-
cause the new specimens had to be made from the materials of a dif-
3. Experimental investigation ferent heat. The yield strength was then estimated as 724 MPa [105 ksi]
corresponding to a 0.2 percent residual strain. The measured ultimate
3.1. Group 1 specimen design strength was 914 MPa [132.6 ksi]. The measured strengths are close to
the typical values reported by the manufacturer.
Ready-mixed concrete with a targeted strength of 27.6 MPa
A total of 24 single anchors placed in six concrete blocks, as illu-
strated in Fig. 2, were first tested, referred as Group 1 tests hereafter, to [4000 psi] was used. Tests at 28 days showed an average strength of
26.2 MPa [3800 psi] using three 100 × 200 mm [4 × 8 in.] cylinders.
evaluate the code-conforming anchor reinforcement. The cast-in an-
chors consisted of a 25 mm [1 in.] diameter ASTM A193 Grade B7 rod Additional cylinder tests during the anchor tests (roughly 58 days)
showed an average compressive strength of 37.2 MPa [5400 psi]. Note
with a heavy hex nut. Plate washer was not used in this group of spe-
cimens because pullout failure was initially not expected to control the that this strength is higher than the design assumption; hence the ad-
verse impact of the under-designed anchor head was greatly reduced.
behavior. All anchors were embedded 152 mm [6 in.] in concrete while
two edge distances were used: 152 mm [6 in.] and 254 mm [10 in.].
Although such headed anchors usually have a larger embedment depth 3.3. Test setup
in practice, this small embedment was selected in order to examine the
effectiveness of anchor reinforcement in an adverse condition. The The loading frame, actuator placement, and instrumentation setup
block length and anchor spacing were determined assuming a concrete used for Group 1 tests are illustrated in Fig. 4. Similar to most previous

      
     
     
     

Fig. 2. Specimens for Group 1 tests (1 in. = 25.4 mm).

111
D. Petersen et al. Engineering Structures 164 (2018) 109–118

Table 1
Summary of reinforced anchor tests in tension.

Specimen ID Reinf. type Edge distance ca1 Load type Concrete strength fc Breakout capacity Peak load Peak disp. Bearing stress Specimen name in Ref.
(in.) (psi) (kips) (kips) (in.) (×fc) [25]

Group 1 specimens
UR10-m – 10 M 5400 32.40 36.56 0.05 – 5132011-A1
UR10-c – 10 C 5400 32.40 35.22 0.10 – 5162011-A2
UR06-m – 6 M 5400 24.30 30.17 0.11 – 5162011-A3
UR06-c – 6 C 5400 24.30 33.66 0.12 – 5162011-A4
RB10-m B (Fig. 2) 10 M 5400 32.40 49.65 0.12 6.13 5172011-B1
RB10-c1 B 10 C 5400 32.40 43.58 0.08 5.38 5172011-B2
RB10-c2 B 10 C 5400 32.40 44.53 0.14 5.49 5202011-C1
RB10-c3 B 10 C 5400 32.40 50.11 0.20 6.18 5202011-C2
RC06-m C (Fig. 2) 6 M 5400 24.30 63.00 0.34 7.77 5182011-B3
RC06-c1 C 6 C 5400 24.30 62.40 0.24 7.70 5182011-B4
RC06-c2 C 6 C 5400 24.30 67.31 0.74 8.30 5232011-C3
RC06-c3 C 6 C 5400 24.30 69.99 0.50 8.63 5232011-C4
RD10-m D (Fig. 2) 10 M 5400 32.40 49.07 0.19 6.05 5252011-D1
RD10-c1 D 10 C 5400 32.40 50.96 0.14 6.29 5252011-D2
RD10-c2 D 10 C 5400 32.40 52.63 0.18 6.49 5272011-E1
RD10-c3 D 10 C 5400 32.40 49.69 0.11 6.13 5272011-E2
RE06-m1 E (Fig. 2) 6 M 5400 24.30 48.24 0.09 5.95 4252011-D4
RE06-m2 E 6 M 5400 24.30 52.28 0.08 6.45 5252011-D3
RE06-c1 E 6 C 5400 24.30 53.01 0.18 6.54 5262011-E3
RE06-c2 E 6 C 5400 24.30 54.90 0.12 6.77 5252011-E4
RF06-m F (Fig. 2) 6 M 5400 24.30 57.46 0.21 7.09 5282011-F1
RF06-c1 F 6 M 5400 24.30 56.38 0.50 6.96 5282011-F4
RF06-c2 F 6 C 5400 24.30 53.65 0.18 6.62 5282011-F2
RF06-c3 F 6 C 5400 24.30 55.05 0.27 6.79 5282011-F3

Group 2 specimens
RG10-m G (Fig. 7) 10 M 6900 36.62 79.91 0.41 3.62 5122012-G1
RG10-c1 G 10 C* 6900 36.62 80.85 NA 3.66 5152012-G2
RG10-c2 G 10 C* 6900 36.62 80.86 0.45 3.66 5152012-G3
RG10-c3 G 10 C* 6900 36.62 80.54 0.49 3.65 5162012-G4

Note: M: Monotonic loading; C: Cyclic loading in displacement control; C*: cyclic loading in load control; 1 kip = 4.45 kN; 1 in. = 25.4 mm; 1 psi = 6.9 kPa.

[0.08 in./min] in the monotonic tests. The cyclic displacement steps for
each 3-cycle group were chosen as 2, 4 (failure for typical unreinforced
anchors), 8, 16, and 32 mm [0.08, 0.16, 0.32, 0.64, and 1.28 in.]. The
loading rate for displacement cycles at or below 8 mm [0.32 in.] were
kept at 2 mm/min [0.08 in./min], same as that in the monotonic tests,
while the load rate was increased to 10 mm/min [0.4 in./min] for the
last two cycle groups cycles in order to reduce test time. The impact of
the higher loading rate was found negligible [23].

4. Discussion of group one tests

The monotonic tests of reinforced anchors are compared with the


Fig. 3. Stress vs. strain relationship of anchor steel (ASTM A193 Grade B7).
tests of unreinforced anchors (shown in dark solid lines) in Fig. 5. In
general, the capacities of reinforced anchors are 34 to 100 percent
anchor tests in the literature, a self-balanced load frame was used. higher than that of unreinforced anchors though the desired steel
Different from other reaction systems, the reaction beams in this test fracture was not achieved. The design of anchor reinforcement in this
setup had small lateral stiffness, thus providing negligible lateral re- test group focused on the number of reinforcing bars and their positions
straint for crack development in concrete during a test. The reactions as required by ACI 318-08 Appendix D. However, concrete developed
were provided 356 mm [14 in.] (2.3hef) from the test anchor to mini- cracks in multiple directions as shown in Fig. 6. Following the cracking,
mize the impact. The tensile force in the test anchor thus would transfer the concrete above the anchor heads lost its confinement, and crushed,
to the reinforcement, before reaching the reaction points. The load resulting in pullout failure in most tests. The maximum concrete
transfer from the closed stirrups to the reaction points may be visua- bearing stress, calculated by the observed ultimate loads divided by the
lized using a strut-and-tie model, similar to the tests by Baba et al. [9]. bearing area ( Abrg ) and the measured concrete strength as listed in
An MTS Model 244.41, 489-kN [110-kip] actuator was used to Table 1, was as low as 5.4fc . This indicates that anchors with code-
apply tension loading to the anchor bolt through a loading plate. One conforming head sizes, which usually are proportioned assuming a
monotonic test and three cyclic tests were conducted for each type of concrete bearing strength of 8.0fc′, might still have experienced the
reinforced anchor specimens, as shown in Table 1, expect for the group pullout failure.
of four anchors with Type E reinforcement. The cracks formed during Anchor bolts in plain concrete had a brittle failure after the concrete
the test of RE06-m1 extended to the region near RE06-m2, which could breakout as shown in Fig. 5. The measured peak load was 163 kN
impact the cyclic behavior and provide false comparison. Hence RE06- [36.6 kips] for the anchor with an edge distance of 254 mm [10 in.].
m2 was tested with monotonic loading. This is higher than the code predicted capacity (144 kN [32.4 kips])
The displacement-controlled loading rates were kept at 2 mm/min calculated with the actual concrete strength, but close to the average
breakout capacity (161 kN [36.3 kips]) from the tests of headed

112
D. Petersen et al. Engineering Structures 164 (2018) 109–118

a) b)

Fig. 4. Experimental test setup for Group 1 Specimens. (a) Schematic; (b) picture.

such that the front concrete was pushed out right after a splitting crack
(a) was observed at about 169 kN [38 kips]. In addition, the corner bars
were actually 76 mm [3 in.] below the top surface because of the large
b)
bending radius of No. 6 bars. These factors contributed to the poor
confinement for the concrete above the anchor head. As a result, a small
breakout cone (76 mm [3 in.] deep) formed mostly above the corner
bars as shown in Fig. 6b, leading to the sudden capacity drop shown in
Fig. 6a.
The corner bars were better distributed in Specimen RD10-m with
Type D anchor reinforcement, in which the anchor bolt was reinforced
by four No. 6 bars from four stirrups. The two wider stirrups were next
to the anchor bolt as shown in Fig. 2, and extended towards the front
face of the concrete block. The distributed bars along the top surface
provided better restraints for concrete as shown in Fig. 6c. As a result,
the load capacity was maintained for a large displacement as shown in
(b) dashed lines in Fig. 5a. On the other hand, the ultimate capacity
(controlled by pullout failure) was similar to that of Specimen RB10-m
because the left side concrete was pushed away after a crack formed
outside the No. 6 stirrup (Fig. 6c) at about 200 kN [45 kips], and the
confinement condition for the concrete near the anchor was thus partly
destructed.
The concrete was better confined in Specimen RC06-m with Type C
reinforcement. Four No. 4 stirrups were extended to the front face with
a cover of 38 mm [1.5 in.], and the corner bars were close to the anchor
along the top surface as shown in Fig. 2. Unlike the specimen with Type
D reinforcement, smaller size stirrups were distributed within a wider
range, providing better confinement for the concrete. The specimen
achieved a load of 280 kN (63 kips), above the yield capacity of the
anchor bolt, and the load capacity was maintained over a large dis-
Fig. 5. Monotonic tension test results of reinforced anchors. (a) ca1 = 254 mm [10 in.]; placement (i.e., 15 mm [0.6 in.]) as shown in dotted lines in Fig. 5b. In
(b) ca1 = 152 mm [6 in.]. addition to better confinement, the four bars placed near the anchor
bolt may have provided better load transfer. Although pullout failure
controlled the test as shown in Fig. 6d, the calculated concrete bearing
anchors/studs in the literature [6]. The code prediction becomes
strength was around 7.7fc. A larger anchor head may increase the
108 kN [24.3 kips] for the unreinforced anchor with an edge distance of
pullout capacity beyond that needed to rupture the anchor bolt, which
152 mm [6 in.], which is a bit lower than the measured capacity 134 kN
was the expected behavior of reinforced anchors.
[30.2 kips], as shown in Fig. 5. Instead of forming an idealized breakout
Specimen RE06-m1 with Type E anchor reinforcement was similar
cone, the concrete developed radial splitting cracks, and the cracks
to Specimen RC06-m except that the legs of four narrow stirrups, as
propagated below the reaction plates, as shown in Fig. 6a.
anchor tension reinforcement, were placed farther to provide sym-
Anchor bolts reinforced with two No. 6 stirrups (Specimen RB10-m)
metric tensile resistance as suggested by existing design guidelines [7].
had brittle failure as shown in dotted lines in Fig. 5a. Before the sudden
However, the behavior of Specimens RE06-m1 was surprisingly dif-
capacity drop near 4 mm [0.16 in.], the test anchor experienced a 2-mm
ferent from that of Specimen RC06-m as shown in Fig. 5b. Concrete
[0.08-in.] slip. This is similar to the observations by Shahrooz et al. [22]
crack initiated above the anchor head and propagated in two different
in their tests of anchor groups in concrete walls without closely spaced
angles: the crack towards the front face was flat similar to that observed
stirrups. The two narrow stirrups did not provide effective restraints
in the test of unreinforced anchors, while the crack towards the back

113
D. Petersen et al. Engineering Structures 164 (2018) 109–118

a) A1-unreinforced b) B1-2#6 (2 legs)

c) D1-4#6’s d) B3-6#4’s+2#3’s

e) D3-6#4’s+2#3’s f) F1-8#4’s
Fig. 6. Typical failure of anchor bolts in Group 1 tests with various patterns of reinforcement.

bent upwards. As a result, the vertical legs of the four larger stirrups designed anchor head was deemed insignificant when taking into
bridged the crack and acted as anchor reinforcement, but the four legs consideration the tests by Hasselwander et al. [15], which included a
on the back was bypassed by the crack as shown in Fig. 6e. The crack on series of tension tests for 44-mm [1.75-in.] high-strength anchor bolts.
the back caused a sudden capacity drop as shown in dashed lines in The anchor heads were made of 102 mm [4 in.] diameter washers,
Fig. 5b. The load capacity was not completely lost because the breakout which was overdesigned corresponding to a maximum bearing stress of
cone was held back by the net of reinforcing bars on the top surface and 6.3fc′. However, the anchors had an edge distances ranging from 25 mm
restrained by the four vertical bars in the front. [1 in.] to 152 mm [6 in.], and no closed ties were provided near the
The importance of closely spaced bars to mobilize farther bars in anchor heads. The failure of the specimens was controlled by side-face
transferring the tensile force from the anchor was demonstrated by blowout with the calculated maximum bearing stresses ranging from
Specimen RF06-m. The anchor bolt was reinforced by eight No. 4 bars 2.7fc′ to 5.3fc′, and the edge distance was found critical to the anchor
from eight stirrups with a spacing of 51 mm [2 in.] as shown in Fig. 2. If capacities.
a strut-and-tie model is used to explain the load transfer, the strut be-
tween the anchor head and the node formed by the outmost stirrup and
the corner bar had an angle of 41° with respect to the horizontal axis, 5. Proposed anchor tension reinforcement
indicating that all the bars were effective. Unfortunately, side-face
blowout occurred, and the concrete above the anchor head crushed at a The code-recommended anchor reinforcement [5,7] can be effective
load of 222 kN [50 kips]. The corresponding bearing strength was found for anchors placed far from any edges such that splitting cracks would
to be 7.1fc . This A breakout crack was identified as shown by the marked not weaken the confinement to the concrete near the anchor bolts. An
lines in Fig. 6f. The crack was almost flat within the range of anchor alternative design method for anchor tension reinforcement was pro-
reinforcement, and bent up towards the reaction plates on the top posed based on the observations made both in Group 1 tests and those
surface. This observation indicates that closely spaced bars can change in the literature. The proposed anchor tension reinforcement, as illu-
the direction of crack propagation, which was also observed in the tests strated in Fig. 7, includes: (1) load-carrying reinforcement in terms of
of reinforced anchors in shear [24]. closed stirrups; (2) crack-controlling reinforcement in all directions that
The observed pullout failure was also observed in the cyclic tests, has a limited edge distance; and (3) local confining reinforcement near
and the detailed behavioral data can be found elsewhere [25]. Com- the anchor head if side-face blowout may control the failure. In addi-
pared with the lack of proper confinement, the impact of under- tion, the tensile load on the anchors eventually needs to be transferred
to the rest of the structure. Sufficient reinforcement, such as stirrups

114
D. Petersen et al. Engineering Structures 164 (2018) 109–118

with the struts ranging from the anchor head to the outmost load-car-
rying stirrups. The crack-controlling reinforcement can thus be de-
termined considering the load transfer from the load-carrying re-
inforcement to the rest of the structure, as shown in Fig. 7. All crack-
controlling reinforcement should be proportioned using a steel stress of
0.6f y , and implemented using small diameter bars evenly distributed
with a small and practical spacing in two orthogonal directions [7]. The
crack-controlling reinforcement should also be properly developed.
In addition to determining the needed reinforcement, design using
STMs should also include capacity checks for struts and nodes. Such
capacity checks require the geometry of struts and nodes, which can be
difficult to define in the design of anchor reinforcement. For example,
all struts in the 3-D space start from the anchor head, which defines the
size of the nodal zone. If the code-specified effective compressive
strength (i.e., 0.85fc′ in ACI 318-08) is used in the capacity check of the
node, the design may result in a large head size. On the other hand,
such large head size is not necessary if anchor pull-out failure is to
control the design, in which a very different concrete strength (i.e., 8.0fc′
in ACI 318-08) is used. Concrete confinement is critical in this case to
ensure proper load transfer. In addition, distributed reinforcement is
required to restrain struts from splitting cracks in typical design using
STMs; however, such measures have been overlooked in the existing
anchor design recommendations, as illustrated in Fig. 1.
The closest four stirrups, as the load-carrying reinforcement, can be
encased by three or four closed hoops in the transverse direction near
the anchor head if the anchors are close to a free side surface. This is to
Fig. 7. Proposed anchor tension reinforcement layout. better confine the concrete that is subjected to large compressive
stresses above the head. These hoops may also help prevent the side
and surface bars, is needed to ensure proper load transfer. face blowout failure observed in this study and that by Hasselwander
The load-carrying reinforcement should be proportioned to carry a et al. [15]. Further study is needed to quantify this part of the design
force equal to the anchor steel capacity in tension. The nominal yield procedure.
strength of steel should be used in the calculation. The load-carrying
reinforcement should be implemented using small-diameter closed 6. Verification of proposed anchor tension reinforcement
stirrups. The stirrup width should be small yet satisfy the bending ra-
dius requirements. The depth of the stirrups should be large enough as 6.1. Group 2 experiment design
shown in Fig. 7 such that the vertical legs are fully developed for the
tension load at both sides of the anchor head. Two stirrups should be Four tests were conducted to verify the proposed anchor tension
placed next to the anchor shaft, where the crack in concrete may in- reinforcement. The specimen design is shown in Fig. 8. The cast-in
itiate above the anchor head under a tensile load. Rather than placing anchors were made of the same 1-in. diameter ASTM A193 Grade B7
all bars within a small distance from the anchor (e.g. 0.5hef) [5,7], rod as Group 1 specimens while the anchor head and the anchor re-
closely spaced stirrups with a center-on-center spacing of 51 mm [2 in.] inforcement were redesigned. The head size was determined assuming
to 76 mm [3 in.] can extend further from the anchor bolt (e.g., 1.0hef). cracks existed in the concrete and the maximum concrete bearing
Although the development length requirements for the anchor re- strength was assumed 7.0fc′, where a nominal design concrete strength
inforcement can be satisfied through the interaction between the closed of 27.6 MPa [4 ksi] was used. The 1-in. diameter anchor could carry a
stirrups and longitudinal bars at all four corners [25], it is re- maximum tensile load of 358 kN [80.4 kips]; thus a net bearing area of
commended that the closed stirrups are at least 8.0ds (where ds is the 18.5 cm2 [2.9 in.2] was needed, and a 3/8-in. thick plate washer
diameter of the stirrups) from the anchor head at both sides [19], as (5.1 × 5.1 cm [2 × 2 in.]) was tack welded to the heavy hex nut, pro-
shown in Fig. 7. viding a net bearing area of 20.6 cm2 [3.2 in.2]. The anchors were again
The crack-controlling reinforcement is needed to cross all possible embedded 152 mm [6 in.] into concrete as illustrated in Fig. 9. The
splitting cracks. The load transfer from the anchor head to the load- concrete beam width was 508 mm [20 in.] and the height was 432 mm
carrying reinforcement can be represented by a strut-and-tie model [17 in.]. The closest tie-down was 305 mm [12 in.] (2hef ) away from the
test anchor.

Fig. 8. Specimen design for Group 2 tests (1 in. = 25.4 mm).

115
D. Petersen et al. Engineering Structures 164 (2018) 109–118

a)

b)

Fig. 9. Strut-and-Tie models for the specimen design for Group 2 tests.

The loading-carrying reinforcement was proportioned to carry a


load of 358 kN [80.4 kips]; thus the required anchor reinforcement
(Grade 60 steel) was found to be 8.6 cm2 [1.34 in.2], which was pro-
vided using eight No. 4 bars from four closed stirrups. Two stirrups
were placed by the anchor shaft and the other two were 51 mm [2 in.]
away. These four stirrups were 356 mm [14 in.] high measured from the Fig. 10. Test setup for Group 2 tests. (a) Schematics; (b) picture.
outside edges. The development length below the anchor head was
254 mm [10 in.] to facilitate the fabrication of the reinforcing cage. The
degree bend on both ends. The closed stirrups and side bars also formed
load transfer from the anchor to the loading-carrying reinforcement
a cage to prevent concrete side face blowout failure; hence, the re-
may cause splitting cracks in two directions, as shown by the strut-and-
commended local confining reinforcement, in terms of closed hoops
tie models in Fig. 9. The splitting force at the location of the horizontal
perpendicular to the anchor shaft near the anchor head shown in Fig. 7,
leg of the stirrups was roughly 60 percent of the tensile force in the
was not provided.
vertical legs of the stirrups; thus the provided reinforcement (4 No. 4’s)
Ready-mixed concrete with a targeted strength of 27.6 MPa
was sufficient. Similarly, the provided reinforcement (2 No. 5’s) were
[4000 psi] was used. Tests at 28 days showed an average strength of
deemed sufficient in the perpendicular direction. The concrete stress at
39.9 MPa [5800 psi] using three 100 × 200 mm [4 × 8 in.] cylinders.
the nodal zone above the anchor head was deemed sufficient because
Additional cylinder tests during the anchor tests (roughly 85 days)
the nodal zone was confined in multiple directions.
showed an average concrete strength of 47.6 MPa [6900 psi].
The rest crack-controlling reinforcement was proportioned using the
The concrete beam was fixed to the strong wall of the UWM
second strut-and-tie model shown in Fig. 9. The direction of the beam
Structures Laboratory at the two tie-down points, as illustrated in
was the main load transfer path. From the overall equilibrium of the
Fig. 10. Two MTS Model 244.31, 245-kN [55-kip] actuators were used
beam, which was loaded at an anchor and simply supported at two tie-
to apply tension loading to the anchor bolt through a loading beam.
down locations, the tie-down closer to the anchor would carry a load of
One displacement-controlled monotonic test and three load-controlled
238.4 kN [53.6 kips]. The tensile force at the top tie was thus 216.5 kN
cyclic tests were conducted, as indicated in Table 1. The loading steps
[48.7 kips], and the required reinforcement was calculated as 8.7 cm2
were chosen as 133 kN [30 kips] (elastic behavior), 248 kN [56 kips]
[1.4 in.2] using a steel stress of 0.6fy. Six No. 5 longitudinal bars were
(first bar yielding according to the monotonic behavior), 320 kN
used on the top face, and the same number of bars were used on the
[72 kips] (90 percent of the peak load), and 378 kN [76 kips] (95 per-
bottom face to facilitate the construction. A similar analysis was con-
cent of the peak load). Three loading cycles were used for each loading
ducted in the transverse directions, and the analysis indicate that No. 3
step, and the specimen was loaded monotonically to failure after the
stirrups with a spacing of 76 mm [3 in.] throughout the beam was
cyclic loading. The loading rate for the monotonic test was kept at
sufficient to restrain the splitting cracks. The dimensions of the nodal
2 mm/min [0.08 in./min] while the loading rate for the cyclic tests was
zone at the bottom of the beam, shown by a shaded rectangular in
0.89 kN/min [200 lbs/min].
Fig. 9, were defined by the load-carrying reinforcement (4 No. 4 stir-
rups), resulting in a nodal area of 152 × 152 mm [6 × 6 in.]. This nodal
zone was sufficient according to the regulations in ACI 318-08 Ap-
6.2. Observed anchor behavior
pendix A [7]. The strut from the bottom node to the tie-down at the top
of the beam can be a 3D bottle-shaped strut. According to RA3.3 of ACI
The tensile behavior of selected anchors with the proposed anchor
318-08, such struts should be crossed by reinforcement to resist the
reinforcement is shown in Fig. 11a, and the measured capacities listed
splitting tensile force from the compression force spreading. Two
in Table 1. The desired ductile steel failure was achieved in all four
longitudinal bars (No. 5) was placed along the side faces. These bars,
specimens, and the fractured anchors ae shown in the picture in
along with the provided stirrups, were found sufficient to confine the
Fig. 11b. The monotonic behavior, shown in dashed lines in Fig. 11a for
struts according to ACI 318-08 [7]. All longitudinal bars had a 90-
Specimen RG10-m, indicated that the specimen behaved elastically

116
D. Petersen et al. Engineering Structures 164 (2018) 109–118

as crack-controlling reinforcement in the longitudinal direction may not


be practical because the concrete members are expected to develop
inelastic behavior in an earthquake. Further tests of headed anchors
a) installed in the plastic hinge zone of reinforced concrete columns in-
dicated that local crack-controlling reinforcement can be effective [26].
The cyclic behavior of Specimen RG10-c2 was reliable, similar to
the monotonic behavior. The initial slope of Specimen G3 was higher
than that in Specimen G1, as shown in Fig. 11a. Before yielding, the
cyclic loading at 248 kN [56 kips] caused insignificant damage such
that the anchor displacements were almost the same at three peak
loads. Under larger loads, the concrete damage caused the anchor dis-
placement to noticeably increase at every consecutive peak load.

6.3. Observed anchor reinforcement behavior

b) Strain gages were placed on four legs of the stirrups, which served
as the load-carrying reinforcement. The positions of the strain gages are
shown by the inserted schematic in Fig. 12. An idealized breakout crack
was used as the reference, and the gages were placed 13 mm [0.5 in.]
below the assumed breakout cracks. Two gages, placed on the stirrups
next to the anchor, had a horizontal position of 19 mm [0.75 in.] from
the center of the anchor, as marked in Fig. 12; however the absolute
position of the stirrups legs was 79 mm [3.1 in.] from the center of the
anchor because the stirrups were 152 mm [6 in.] wide. Similarly, the
gages on the two outer stirrups were marked at 70 mm [2.75 in.] in
Fig. 12, and the absolute position of the stirrups was 103 mm [4.1 in.]
from the center of the test anchor.
The gages on the closer bars initially recorded larger strains than
those on the farther bars, which is similar to the observation by Lee
et al. [18]. In addition, the measured strains indicated that all No. 4
Fig. 11. Behavior of reinforced anchors with proposed anchor reinforcement. (a) Load- stirrup legs, as the load-carrying reinforcement, yielded when the an-
displacement behavior; (b) failure pictures. chor bolt fractured. Strains in the crack-controlling reinforcement were
not monitored. Further tests may be needed to verify the effectiveness
before the anchor yielded at a displacement of 2 mm [0.08 in.]. The of using strut-and-tie models in the anchor reinforcement design.
measured anchor displacement, within this range of behavior, was
largely from the concrete compressive deformation above the head in 7. Summary and conclusions
addition to the elongation of the anchor shaft. The concrete at the top
surface developed splitting cracks as shown in Fig. 11b, and the split- It is well recognized that the concrete around anchor connections
ting cracks were restrained by the crack-controlling reinforcement. needs proper protection. The existing anchor reinforcement re-
The post-yielding deformation of the anchor was mainly from the commended in the literature, mainly U-shaped hanger steel, was eval-
inelastic deformation of the anchor shaft. The total elongation of the uated using experimental tests in this study. Five patterns of anchor
152-mm [6-in.] long anchor shaft, with a peak strain of 5 percent tension reinforcement were implemented for single cast-in anchors,
(Fig. 4), was calculated as about 8 mm [0.3 in.], which is similar to the made of 1-in. diameter ASTM A193 Grade B7 rods with a heavy hex
measured anchor displacement at the peak load, as shown in Fig. 11a. A nut. The anchor reinforcement, in terms of the legs of closed stirrups,
sudden increase in the anchor displacement was captured at about was proportioned to carry a force equal to the tensile capacity of the
320 kN [72 kips], which was attributed to the flexural crack in the anchors, which were expected to develop ductile steel fracture.
concrete beam at the anchor location (Fig. 11b). The flexural crack was Compared with the anchors embedded in plain concrete, a capacity
again restrained upon further loading by the longitudinal reinforcement increase ranging from 40% to 130% was observed, which is similar to
in the beam, specifically designed for crack controlling purposes for the observations in the literature. However, the desired ductile steel
these specimens. Note that using longitudinal bars in flexural members fracture was not achieved. Splitting cracks developed in concrete, and
the concrete around the anchor head lost its confinement and crushed

a) under monotonic loading; b) under cyclic loading


Fig. 12. Measured strains in load-carrying reinforcement.

117
D. Petersen et al. Engineering Structures 164 (2018) 109–118

prematurely, resulting in anchor pullout failure. [2] Asia-Pacific Economic Cooperation. Earthquake disaster management of energy
Based on these observations and other tests in the literature, re- supply system of APEC member economies. Taipei, China: Energy Commission,
Ministry of Economic Affairs; 2002.
commendations for anchor tension reinforcement were proposed. The [3] American Concrete Institute (ACI). State-of-the-art report on anchorage to concrete,
proposed anchor tension reinforcement includes: (1) load-carrying re- ACI 355.1R-91 (reapproved in 1997), American Concrete Institute, Farmington
inforcement in terms of closed stirrups; (2) crack-controlling re- Hills. MI.
[4] Comité Euro-International du Béton (CEB). Fastenings to concrete and masonry
inforcement in all directions; and (3) local confining reinforcement near structures: state of the art report. Thomas Telford Service Ltd., London; 1997.
the anchor head if side-face blowout may control the failure. [5] Federation Internationale du Beton (fib). Fastenings to concrete and masonry
Specifically, the load-carrying reinforcement is recommended to be structures. Special Activity Groups (SAG) 4 draft report, 2008, Obtained from Dr.
Eligehausen.
proportioned to carry a force equal to the design tensile capacity of the [6] Petersen D. Seismic behavior and design of cast-in-place anchors MS Thesis
anchors, and implemented using small-diameter closed stirrups. The Milwaukee, WI: University of Wisconsin; 2011.
closely spaced stirrups with a center-on-center spacing of 51 mm [2 in.] [7] American Concrete Institute (ACI). Building code requirements for structural con-
crete (ACI 318-08). Farmington Hills, MI; 2008.
to 76 mm [3 in.] can extend up to 1.0hef from the anchor bolt. Crack-
[8] European Committee for Standardization. Design of fastenings for use in concrete –
controlling reinforcement are necessary crossing all possible cracks. The Part 4–2: headed fasteners. CEN/TS 1992-4-2, Brussels Belgium; 2009.
design of the crack-controlling reinforcement may use two strut-and-tie [9] Industry Handbook Committee. PCI design handbook: precast and prestressed
models: one describing the load transfer from the anchor head to the concrete, 6th ed. MNL-120-04. Chicago, IL: PCI; 2004.
[10] Anderson N, Meinheit D. A review of headed-stud design criteria in the sixth edition
load-carrying reinforcement, and the other describing the load transfer of the PCI design handbook. PCI J 2007;52(1):82–100.
from the load-carrying reinforcement to the rest of the structure. All [11] Sharma A, Eligehausen R, Asmus J. Comprehensive analytical model for anchorages
crack-controlling reinforcement should be proportioned using a steel with supplementary reinforcement. In: The 3rd international symposium on con-
nections between steel and concrete. Stuttgart, Germany; 2017.
stress of 0.6f y , and implemented using small diameter bars evenly dis- [12] Berger W. Load-displacement behavior and design of anchorages with headed studs
tributed with a small and practical spacing in two orthogonal direc- with and without supplementary reinforcement under tension load (PhD Thesis)
tions. In addition, the anchors and the load-carrying reinforcement NY: Institute of Construction Materials, University of Stuttgart; 2015. [in German].
[13] Cannon R, Godfrey D, Moreadith F. Guide to the design of anchor bolts and other
close to the anchor can be encased by three or four closed hoops in the steel embedments. Concr Int 1981:28–41.
transverse direction near the anchor head if the anchors are close to a [14] Shipp J, Haninger E. Design of headed anchor bolts. Eng J Am Inst Steel Constr
free side surface. 1983;20:58–69.
[15] Hasselwander G, Jirsa J, Breen J, Lo K. Strength and behavior of anchor bolts
The proposed anchor reinforcement was evaluated with additional
embedded near edges of concrete piers Research Report 29-2F TX: Center for
four tests. Splitting cracks initiated during the tests and were success- Transportation Research, the University of Texas at Austin; 1974.
fully restrained by the crack-controlling reinforcement. Reinforced with [16] Widianto O, Patel C. Design of anchor reinforcement in concrete pedestals. In:
the proposed anchor reinforcement, all four single anchors made with Proceedings of the 2010 structures congress. Orlando, FL; 2010. p. 2500–11.
[17] Baba N, Kanai S, Nishimura Y. Experimental study on stress transfer of joints
1-in. diameter ASTM A193 Grade B7 rods, developed their full tensile connected steel member with reinforced concrete member using anchor bolt. In:
capacity and ductile behavior with an embedment depth of six times the Proceedings of the 14th world conference on earthquake engineering. Beijing,
anchor diameter. China; 2008.
[18] Lee N, Kim K, Bang C, Park K. Tensile-headed anchors with large diameter and deep
embedment in concrete. ACI Struct J 2007;104(4):479–86.
Acknowledgments [19] Sharma A, Eligehausen R, Asmus J. Comprehensive experimental investigations on
anchorages with supplementary reinforcement. In: The 3rd international sympo-
sium on connections between steel and concrete. Stuttgart, Germany; 2017.
The study reported in this paper is from a project funded by the [20] Saari W, Hajjar J, Schultz A, Shield C. Behavior of shear studs in steel frames with
National Science Foundation (NSF) of United States under Grant No. reinforced concrete infill walls. J Constr Steel Res 2004;60(10):1453–80.
0724097. The authors gratefully acknowledge the support of Dr. Joy [21] Baran E, Schultz A, French C. Tension tests on cast-in-place inserts: influence of
reinforcement and prestress. PCI J 2006;51(5):88–108.
Pauschke, who served as the program director for this grant. The au- [22] Shahrooz B, Deason J, Tunc G. Outrigger beam-wall connections. I: Component
thors also thank the colleagues in ACI committee 355 for their valuable testing and development of design model. J Struct Eng 2004;130(2):253–61.
inputs. Any opinions, findings, and recommendations or conclusions [23] Shahrooz B, Tunc G, Deason J. Outrigger beam-wall connections. II: subassembly
testing and further modeling enhancements. J Struct Eng 2004;130(2):262–70.
expressed in this material are those of the authors and do not ne-
[24] Petersen D, Zhao J. Design of anchor reinforcement for seismic shear loads. ACI
cessarily reflect the views of NSF. Struct J 2013;110(1):53–62.
[25] Petersen D, Lin Z, Zhao J. Reinforcement for headed anchors Final report (volume
References II) for project – behavior and design of cast-in-place anchors under simulated
seismic loading. Milwaukee, WI: University of Wisconsin; 2013.
[26] Lin Z, Luke B, Shahrooz B, Zhao J. Headed anchors in plastic hinge zone of re-
[1] Lifeline Earthquake Engineering (ASCE). Northridge earthquake: lifeline perfor- inforced concrete members Final report (volume III) for project – behavior and
mance and post-earthquake response. A report to U.S. department of commerce; design of cast-in-place anchors under simulated seismic loading. Milwaukee, WI:
NIST Building and Fire Research Laboratory. Gaithersburg, MD; 1997. University of Wisconsin; 2013.

118

You might also like