You are on page 1of 36

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

net/publication/269122541

Allowable Tensile Stress Limit at Prestress Transfer

Conference Paper · April 2009


DOI: 10.1061/41031(341)167

CITATIONS READS

0 872

2 authors, including:

Robin Tuchscherer
Northern Arizona University
23 PUBLICATIONS   127 CITATIONS   

SEE PROFILE

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

Arizona Tri-University Transportation Research Collaboration View project

All content following this page was uploaded by Robin Tuchscherer on 30 July 2019.

The user has requested enhancement of the downloaded file.


1 TENSILE STRESS LIMIT FOR PRESTRESSED CONCRETE AT

2 RELEASE: ACI 318-08

3 Robin G. Tuchscherer and Oguzhan Bayrak

4 ACI member Robin Tuchscherer is a PhD candidate at the University of Texas at Austin.

5 He received his MS from the University of Texas at Austin. His research interests include the

6 behavior of reinforced and prestressed concrete.

7 ACI member Oguzhan Bayrak is an Associate Professor of civil, architectural, and

8 environmental engineering and a Fellow of the Clyde E. Lee Endowed Professorship at the

9 University of Texas at Austin. He is a member of ACI-ASCE Committee 441, Reinforced

10 Concrete Columns; ACI Committees 341, Earthquake-Resistant Concrete Bridges; E803,

11 Faculty Network Coordinating Committee; and Joint ACI-ASCE Committee 445, Shear and

12 ABSTRACT

13 The tensile stress limit for prestressed concrete beams at release was evaluated

14 experimentally for seven 54 in. (1370 mm) deep bridge beams. Additionally, an extensive

15 amount of material data was collected through testing and literature review. The purpose of

16 the experimental program was to characterize the mechanical properties of concrete,

17 including the relationship between the in-situ strength of a beam and its representative

18 material strength. Beams fabricated as part of this study cracked with an applied stress less

19 than half of the apparent strength determined from material tests. The ACI 318 tensile stress

20 limit was not sufficient to prevent cracking. It was concluded that limiting the extreme fiber

21 tensile stress to 4 f 'ci (0.33 f ' ci MPa) will adequately prevent cracking at release.

22 Keywords: precast/prestressed beam, concrete tensile strength, cracking at release, strain

23 gradient

1
1 INTRODUCTION

2 This research program was funded by the Texas Department of Transportation

3 (TxDOT) to determine the cause of flexural cracking of prestressed beams at release. Girders

4 that exhibited cracking were 54 in. (1370 mm) deep, relatively short in length (60 feet [18.3

5 m] and less), and contained highly eccentric strand configurations. Cracking was typically

6 observed 4 ft. (1.22 m) from the ends (Fig. 1). Beams that cracked in the field had relatively

7 high top-fiber tensile stresses (as high as 6 f 'ci psi [0.50 f ' ci MPa]). TxDOT engineers,

8 aware that the tensile strength of concrete is conventionally equal to 7.5 f 'ci psi (0.62 f ' ci

9 MPa), were interested in determining the cause of the cracking.

10 Crack control is an important issue for prestressed beam manufacturers. Beams are

11 typically released less than 24 hours after they are batched. Therefore, concrete is relatively

12 green and actively hydrating. Elastic shortening, creep, and shrinkage will relieve some of the

13 applied stress; however, these factors also contribute to the long term camber of the beam. As

14 a beam continues to camber, new cracks can form and existing cracks will propagate. If crack

15 control reinforcement is not provided and cracking occurs, the beam will not pass inspection.

16 It will either be rejected or cracks will be repaired via epoxy injection; an expensive solution.

17 As a result, TxDOT funded the current research program. The objective of the research is to

18 gain a better understanding of the cracking behavior of large-scale prestressed beams. Tensile

19 stress limits for crack control were investigated. Consequently, the research team observed

20 that the ACI 318 §18.4.1 tensile stress limit was not sufficient to prevent cracking at the time

21 of release.

22 To study current tensile stress limits and crack control provisions, the research team

23 fabricated seven full-scale specimens with a test region on each side, resulting in 14 full-scale

24 tests. All of the beams represented the worst-case scenario of girders used in the State of

25 Texas, yet they met the requirements of the ACI code. Current code provisions, concrete

2
1 material properties and typical design assumptions were examined in order to determine their

2 contribution to the cracking problem. Also, the provisions in ACI were evaluated based on

3 historic data and data collected as part of the current research program. As a result,

4 contributing factors to the cracking problem were identified and recommendations made to

5 control cracking with greater assurance.

6 BACKGROUND: ACI 318 PROVISIONS

7 In 1958, ACI-ASCE Committee 3231 presented the first tentative recommendations

8 for prestressed concrete. Their recommendations were adopted by ACI Committee 318 and

9 included in the 1963 edition of the ACI 318 building code2. From 1963 until 1977 the

10 extreme fiber tensile stress before losses due to creep and shrinkage was limited as follows2:

11 Allowable tension stress in members without auxiliary reinforcement (unprestressed

12 or prestressed) in the tension zone:

13 3 f ' ci psi (0.25 f ' ci MPa)

14 Where the calculated tension stress exceeds this value, reinforcement shall be

15 provided to resist the total tension force in the concrete computed on the assumption

16 of an uncracked section.

17 The allowable tensile stress limit remained unchanged until the release of the 1977

18 update to the ACI 318 building code3. After 1977, the tensile stress at the ends of simply

19 supported members was allowed to be as high as 6 f 'ci psi (0.50 f ' ci MPa). The tensile

20 stress at other locations continued to be limited to 3 f 'ci psi (0.25 f ' ci MPa). Since 1977,

21 the following limitation has remained essentially unchanged4:

22 Where computed concrete tensile strength, ft, exceeds 6 f ' ci psi (0.50 f ' ci MPa) at

23 the ends of simply supported members, or 3 f ' ci psi (0.25 f ' ci MPa) at other

24 locations, additional bonded reinforcement shall be provided in the tensile zone to

3
1 resist the total tensile force in concrete computed with the assumption of an

2 uncracked section.

3 Changing the tensile stress at the ends of simply supported members from 3 to 6 f 'ci

4 psi (0.25 to 0.50 f ' ci MPa) is a key issue. However, to the best of our knowledge, research

5 substantiating the change is unavailable. An explanation5 supporting the change is given as

6 follows.

7 1. Most standard prestressed concrete units, especially hollow-core slabs, are cast

8 with straight strands. At the end of the units, where DL is not fully effective, tension

9 occurs in the top fibers which is greater than the current Code value but not sufficient

10 to cause cracking. To place top steel at the ends is an added production expense and

11 many producers have shown that performance is not impaired when it is omitted,

12 according to the relief provided by Section 18.4.3.

13 [18.4.3 – Permissible stresses … shall be permitted to be exceeded if shown by test or

14 analysis that performance will not be impaired.]

15 2. By increasing the permissible stress to a value just below cracking, it becomes

16 more logical when the stress is exceeded to require the addition of bonded auxiliary

17 reinforcement to withstand the tension force and to control crack width and depth.

18 3. Note that this is a temporary stress condition immediately after transfer of

19 prestressing forces and that the loss effects due to creep and shrinkage in the concrete

20 and relaxation in the tendons will always work to relieve these stresses.

21 4. An analytical study on heavily loaded standard prestressed concrete products

22 showed no correlation between low release tension and better shear behavior. The

23 shear stress conditions at the ends of these simply supported units indicate that higher

24 allowable tension at the ends should be permitted.

4
1 BACKGROUND: TENSILE STRENGTH OF CONCRETE

2 Concrete tensile strength can be measured several different ways. Since past attempts

3 to measure the tensile strength uniaxially have yielded inconsistent results, the tensile

4 strength of concrete is not commonly tested directly. An illustration of a direct tension test is

5 shown in Fig. 2(i). The average strength of concrete tested in direct tension is typically taken

6 to be equal to 4 f ' ci psi (0.33 f ' ci MPa)6-8. More commonly, the tensile strength of

7 concrete is taken as the modulus of rupture (MOR) determined with a bending test (ASTM

8 C789); or as the splitting strength established with a split cylinder test (ASTM C49610). An

9 illustration of a split cylinder and MOR test is shown in Fig. 2 (ii and iii). The average

10 strength of concrete determined from a split cylinder and MOR test is typically taken to be

11 equal to 6 and 7.5 f ' ci psi (0.50, and 0.62 f ' ci MPa) respectively.

12 The primary reason that there is such a varied difference in the strength values

13 associated with the different testing methods can be attributed to the differences in the strain

14 gradients (Fig. 2). The tensile strength of a concrete specimen will decrease as the volume of

15 the concrete that is stressed in tension increases. It follows that as the depth of a MOR

16 specimen increases; the strain gradient will decrease, approach uniformity, and cause the

17 tensile strength to approach a value of 4 f ' ci psi (0.33 f ' ci MPa).

18 In addition to the tensile strain gradient, plastic shrinkage stresses may affect the

19 tensile strength of a specimen. Moist concrete within a specimen restrains the drying concrete

20 at the surface from shrinking thereby inducing tensile stress. The plastic shrinkage stress is

21 added to the flexural-tensile stress resulting in a lower flexural-tensile capacity available to

22 resist cracking.

23 RESEARCH SIGNIFICANCE

24 Some of the specimens tested as part of this experimental program cracked at the time

25 of prestress release even though the applied tensile stress was less than or equal to the

5
1 permissible limits specified in Chapter 18 of ACI 3184. An extensive literature review was

2 conducted in order to determine the basis for the ACI 318 tensile stress limit; along with the

3 empirical expression that relates concrete tensile and compressive strengths. Full-scale beam

4 specimens were fabricated and tested in order to examine typical design assumptions, and the

5 strength differences that exist between in-situ concrete and representative control specimens.

6 Based on the findings of this study, a modification to the current ACI 318 tensile stress limit

7 is proposed.

8 PREVIOUS RESEARCH

9 The relationship between concrete compressive strength and tensile strength is an

10 empirically derived expression. According to ACI 318-05 §9.5.2.34, the flexural tensile

11 strength or MOR of normal weight concrete is related to the compressive strength according

12 to Eq. 1:

13 fr = 7.5 f ' ci … (psi) (1)

14 fr = 0.62 f ' ci … (MPa)

15 In order to better understand the empirical basis of Eq. 1 a thorough literature review

16 of previous MOR research was conducted. Fig. 3 presents a compilation of 1330 data points

17 taken directly from MOR research conducted over the past 80 years11-22. MOR beams tested

18 as part of the current study23 have also been added to the compilation of data illustrated in Fig.

19 3. Included in Fig. 3 are a wide range of variables that influence the tensile strength of

20 concrete. The majority of the data points were from standard 6x6x21 in. (150x150x530 mm)

21 normal-weight concrete beams, loaded at their third points, and cured for 28 days. A small

22 number of the data points are from beams ranging in size between 4x4x14 and 7x10x38 in.

23 (100x100x360 and 180x250x970 mm). Other variables include aggregate size and type,

24 cement type, and curing conditions. Varied aggregate properties include shape, surface

25 texture, and modulus of elasticity16. Aggregate sizes varied from 0.38 to 2.5 in. (10 to 64

6
1 mm)13, 16, 22. The effects of using either Type I or III cement was examined17-19. The different

2 curing conditions that were investigated include varying combinations of moist-curing, dry-

3 curing, or heat-curing11-13, 17, 19. Curing conditions had the most pronounced affect on the

4 large amount of scatter. In general, the average MOR of a heat-cured specimen was

5 approximately 30% less than that of moist-cured specimen19. By and large, the cloud of 1330

6 data points encompasses the full range of tensile strength values that can be expected from a

7 MOR test.

8 Literature examining the effect of specimen depth on the flexural strength of concrete

9 was also encountered. Tucker24 summarized research25- 27


that examined the effect of

10 specimen depth on the MOR of third-point loaded concrete beams as follows: Abrams25

11 showed a 6.5% reduction in MOR for an increase in beam depth from 4 to 10 in. (100 to 250

12 mm); Gonnerman and Shuman26 showed a 6.5% reduction in MOR for an increase in beam

13 depth from 3 to 7.5 in. (75 to 190 mm); and Reagel and Willis27 showed an 11.5% reduction

14 in MOR for an increase in beam depth from 4 to 10 in (100 to 250 mm). Wright28 determined

15 that an increase in specimen depth from 3 to 8 in. (75 to 200 mm) reduced the modulus of

16 rupture by 28% under third-point loading.

17 According to Collins and Mitchell7, the cracking stress is inversely proportional to

18 the fourth root of the depth. Thus, an increase in the depth of a beam from 6 in. (150 mm) to

19 54 in. (1370 mm) decreases the cracking stress from 7.5 to 4.3 f ' ci psi (0.62 to 0.36 f ' ci

20 MPa). In other words, 6( 54 )0.25


⋅ 7 .5 = 4 .3 .

21 According to the Féderation Internationale du Béton (FIB)8, as the depth of the beam

22 increases, the flexural strength approaches the axial tensile strength. Accordingly, they relate

23 the flexural tensile strength to the direct tensile strength according to the following

24 relationship.

7
0.7
⎛h ⎞
1 + α fl ⎜⎜ b ⎟⎟
1 f ct , fl = f ctm ⎝ ho ⎠ (2)
0.7
⎛h ⎞
α fl ⎜⎜ b ⎟⎟
⎝ ho ⎠

2 where fct,fl is the flexural tensile strength of concrete; fctm is the mean axial tensile strength; hb

3 is the beam depth; ho is equal to a standard depth of 100 mm (3.94 in.); and αfl is a coefficient

4 that depends on the brittleness of concrete (varies between 1.0 and 2.0). According to Eq. 2,

5 the flexural tensile strength of a 54 in. (1370 mm) deep member is equal to 4.6 f ' ci psi

6 (0.38 f ' ci MPa) – assuming that the mean axial tensile strength, fctm, is equal to 4 f ' ci psi

7 (0.33 f ' ci MPa) and the coefficient, αfl, is equal to 1.0.

8 EXPERIMENTAL PROGRAM

9 From the onset of the research program, it was theorized that the cracking problem

10 was based on rough assumptions of either the material properties of concrete or the

11 mechanics of prestress transfer. To study this problem, seven full-scale specimens were

12 fabricated and both ends tested. Data attained from the tests was used to verify typical

13 engineering design calculations. Also, additional material data was collected in order to

14 examine the correlation between the in-situ strength and control specimen strengths.

15 The specimen design was determined by replicating the worst-case scenario of actual

16 54 in. (1370 mm) deep bridge girders used in the State of Texas. The mix design,

17 reinforcement, and strand configuration were identical to actual beams in the field. For some

18 of the test specimens, the concrete compressive strength at release was similar to values

19 specified in the field. In other cases, it was greater so that different levels of top fiber tensile

20 stress could be attained. A detailed description of the experimental program follows.

8
1 Specimen Preparation

2 Seven full-scale 54 in. (1370 mm) deep prestressed concrete girders were designed

3 and reinforced in accordance with TxDOT standard details (Fig. 4). It was assumed that the

4 “end region” of a typical girder in the field was 10 ft. (3.05 m) long. Hence, test specimens

5 were designed to be 20 ft. (6.10 m) long so that two simultaneous tests could be conducted on

6 10 ft. (3.05 m) test regions. A 4 ft. (1.22 m) long by 1.5 in. (40 mm) deep block-out was

7 provided at the ends of the beams. The block-out detail is commonly used to accommodate a

8 thickened slab located along an expansion joint. Beams with a block-out contain a stress

9 concentration at the location of discontinuity increasing the likelihood of cracking. Beams in

10 the field exhibited cracking approximately 4 ft. (1.22 m) from their ends whether or not a

11 block-out was present. For example, the beam shown in Fig. 1 cracked at release, yet it did

12 not have a block-out. Based on this information, all test specimens contained the block-out in

13 order to represent the worst-case scenario, and determine whether or not the cause of the

14 cracking problem could be attributed to the block-out. All specimens except Beam 4

15 contained 12 strands located 2 in. (50 mm) from the bottom of the beam (Fig. 5). This highly

16 eccentric strand configuration was considered to be the worst-case scenario used by TxDOT.

17 The purpose of the strand configuration of Beam 4 was to examine the benefits provided by

18 lowering the eccentricity. The location of the raised strands for Beam 4 was selected based on

19 dimensional requirements of the prestressing bed.

20 Each end of each beam had differing amounts of crack control reinforcement. Along

21 the north half of each specimen, 2.4 in2 (1550 mm2) of crack control reinforcement was

22 provided; this represents the amount of reinforcement required to control cracking if the

23 applied tensile stresses exceed 6 f ' ci psi (0.50 f ' ci MPa). Along the south half, 0.6 in2

24 (390 mm2) was provided; this is considered to be a negligible amount of crack control

9
1 reinforcement. Therefore, the south end of each specimen would meet the requirements of

2 ACI §18.4.1 when the applied tensile stress is less than 6 f ' ci psi (0.50 f ' ci MPa).

3 Steel reinforcement was standard Grade 60 deformed bars meeting the requirements

4 of ASTM A61529 (fy = 60 ksi [410 MPa]). The cross sectional area of reinforcement was

5 assumed based on the nominal sizes given in ASTM A615. The modulus of elasticity was

6 29,000 ksi (200 GPa).

7 Prestressing strand was standard low-relaxation Grade 270 0.5 in. (13 mm) meeting

8 the requirements of ASTM A41630 (fpy = 243 ksi [1680 MPa]).

9 Instrumentation

10 Strain gauges were affixed to each strand in order to verify the magnitude of the

11 prestressing force at release. Strain gauges were affixed to the prestressing strands along

12 individual wires. A series of tests were conducted on 3 ft. (910 mm) lengths of strand with

13 gauges attached in order to calibrate the gauge values to the corresponding magnitude of

14 force in the strand.

15 Additionally, instrumented steel reinforcement was located longitudinally along the

16 top and bottom flanges in order to measure the strain profile in each specimen at release. Two

17 #5 (#16) bars in the top flange were instrumented and located 52 in. (1320 mm) from the

18 bottom of the beam. Two #4 (#13) bars in the bottom flange were instrumented and located 6

19 in. (150 mm) from the bottom of the beam (Fig. 6). Strain readings from the instrumented

20 bars were used to determine a straight line strain profile at release for uncracked specimens.

21 The “straightness” of the profile line was validated from additional instrumentation and

22 measurements of Beam 123.

10
1 Stressing of Strands

2 Strands were individually stressed using a single strand hydraulic ram and pump.

3 During stressing operations, the force per strand was verified based on corresponding strain

4 gauge values and a hydraulic pressure gauge connected to the ram.

5 Concrete Casting

6 All specimens were cast using the same mixture design that contained 6.5 sacks of

7 Type III cement per cubic yard of concrete. This concrete mixture proportion is typical of one

8 used by most precast prestressed beam manufacturers for their final batch of the day.

9 Generally, manufacturers will add an extra half to one sack of cement per yard to the

10 afternoon batches so that the specified compressive strength can be achieved first thing the

11 next morning. Characteristics of the mixture proportions are summarized in Table 1. In order

12 to replicate field conditions as closely as possible, the coarse and fine aggregate were

13 supplied by a local precast manufacturer and delivered in a ready-mix truck. At the Ferguson

14 Structural Engineering Laboratory, 90% of the water was added to the aggregates followed

15 by all the cement. The admixtures were diluted into the remaining 10% of water and added

16 last. Finally, the drum of the ready-mix truck was turned approximately 200 times. Pneumatic

17 vibrators attached to the steel formwork and rod vibrators were used to consolidate the

18 concrete in each specimen.

19 Curing

20 All beam specimens were covered with a sheet of plastic while they cured. Beams 3

21 through 7 were covered with both wet burlap and a plastic sheet.

22 Temperature Match Curing

23 Temperature is an important factor in the strength development of concrete, especially

24 within the first 24 hrs. The maturity of concrete is determined by multiplying an interval of

25 time by the temperature of the concrete. Maturity has been shown to be an excellent

11
1 indication of the development of strength of concrete31. In order to better represent the

2 concrete in a 54 in. (1370 mm) deep girder with a standard control specimen it is necessary

3 that the time-temperature profiles match one another.

4 For this study, thermocouples were used to measure temperature values, and through

5 the use of an electronically-controlled match-curing system, cylinders were heated

6 accordingly. It was of interest to the research team to study the difference in temperature

7 within a beam’s cross-section during curing (Fig. 7); including the sensitivity that

8 thermocouple location has on concrete compressive strength. It was determined that the

9 compressive and tensile strength of a match-cured cylinder was unaffected by thermocouple

10 location after the occurrence of the peak beam temperature (12 to 16 hours after batch for this

11 study). For beams fabricated at the Ferguson Laboratory, control specimens were typically

12 cured using a relatively cool location in order to conservatively represent the concrete in the

13 top flange (Fig. 8). For beams that cracked, the temperature of the corresponding control

14 cylinders was less than the concrete in the top flange at the time of release.

15 Cylinders used to determine the compressive strength of concrete were match-cured

16 using an electronically controlled system. Beams used to determine the MOR of concrete

17 were cured adjacent to the 54 inch (1370 mm) girders.

18 Material Testing

19 The concrete compressive strength was determined in the same manner as determined

20 by a precast fabricator. The compressive strength of standard 4 in. (100 mm) cylinders was

21 measured according to ASTM C3932. Cylinders were capped with reusable neoprene pads

22 and steel retainers meeting the requirements of ASTM C123133. Talcum powder was used to

23 break the bond between the cylinder and the neoprene. The cylinders were tested with a

24 stress-controlled, hydraulically actuated, universal testing machine at a rate of approximately

25 35 psi/sec (0.25 MPa/sec).

12
1 Modulus of elasticity (MOE) of concrete was measured in accordance with ASTM

2 C46934. However, the size of the MOE specimen that was tested was not consistent with the

3 requirements of ASTM C469. A 54 in. (1370 mm) tall by 6 in. (150 mm) diameter cylinder

4 was used to determine the modulus of elasticity of a beam’s top flange (Fig. 9). The “tall”

5 cylinder was designed to be the same height as the 54 in. (1370 mm) deep girders in order to

6 accurately represent the segregation that may potentially occur in the top flange. It was not

7 possible to match-cure the tall MOE specimens. Therefore, they were tested once the

8 compressive strength of ambient-cured cylinders was equal to the previously determined

9 compressive strength of the beam specimen at release. On average, the bottom of the tall

10 cylinders had a MOE 3% greater than the top. Such a small difference was considered

11 insignificant given the scatter of data associated with MOE values.

12 The flexural strength or MOR of standard 6x6x21 in. (150x150x530 mm) beams was

13 measured according to ASTM C789.

14 Release

15 A typical precast beam manufacturer will release his beams within 12 to 18 hours

16 from the time the concrete is batched. The only factor that determines when a beam is

17 released is the specified compressive strength of concrete. Once the specified compressive

18 strength is attained, the beam can be released. Typically, the specified compressive strength

19 at release is approximately 4,000 to 5,000 psi (28 to 34 MPa). Specimens fabricated for this

20 project were released when the compressive strength was between 4,500 to 10,700 psi (31 to

21 74 MPa). Strands were released by torch cutting them one at a time. The strands were cut

22 slowly in order to minimize sudden movement of the beams during release. For Beam 3, an

23 8000 lb (3600 kg) concrete block was placed on top of the top flange prior to releasing the

24 strands. The block was used as a dead weight intended to resist the dynamic affects of cutting

13
1 the strands. It was observed that the block had a minimal affect on cracking and crack widths

2 at release. Therefore, subsequent beams were released without using the dead weight.

3 EXPERIMENTAL OBSERVATIONS

4 Key parameters of test specimens are summarized in Table 2. The intent of Beam 1

5 was to calibrate testing procedures and data collection. Beam 2 was released as soon as the

6 compressive strength of concrete was greater than 4000 psi (28 MPa), the same as would

7 typically happen in the field. Beam 2 cracked extensively at release and afterwards. The

8 intent of Beam 3 was to repeat the Beam 2 test with additional measures taken to prevent

9 cracking; such as covering the beam with wet burlap during curing, and placing a concrete

10 block on top of the beam to minimize the dynamic effects of releasing the strands. Regardless,

11 Beam 3 cracked as extensively as Beam 2 at release and afterwards. The intent of Beams 4

12 and 5 was to eliminate cracking at release by either reducing the strand eccentricity or

13 increasing the compressive strength of concrete. After Beam 5 cracked, it was theorized that

14 limiting the top fiber tensile stress to a value less than 4 f ' ci psi (0.33 f ' ci MPa) would

15 eliminate cracking. Beam 6 verified the theory for a beam with a short release time. Beam 7

16 verified the theory for a beam with a higher prestress force.

17 For uncracked sections, the stress applied to the extreme fiber was calculated based on

18 the assumption that planar sections remained planar. This assumption is not applicable in

19 regions where the strain distribution is nonlinear; such as the end region of a prestressed

20 member (referred to as a “D-region”). Typically, in the case of a 54 in. (1370 mm) girder, the

21 D-region is assumed to extend 54 in. (1370 mm) from the end of the beam. The beam

22 specimens fabricated as part of this study were instrumented in order to evaluate the extent of

23 their D-regions; which turned out to be less than expected. There were some discrepancies for

24 planar sections 1 and 2 ft. (0.31 and 0.61 m) from the end of the beam. This was to be

25 expected for a 54 in. (1370 mm) deep member. However, for all test specimens, it was

14
1 possible to consistently predict strain behavior using linear elastic methods with reasonable

2 accuracy at a distance of 3, 4, and 5 ft. (0.91, 1.22, and 1.52 m) from the beam’s end (Fig. 10).

3 Therefore, from the perspective of a design engineer, the stresses predicted at the location of

4 prestress transfer and beyond are reliable and accurate.

5 Even though cracking had been observed 4 ft. (1.22 m) from the end of a beam, the

6 maximum tensile stress was assumed to be located at a distance equal to the theoretical

7 transfer length (ACI §12.9.1) – between 20 and 33 in. (510 and 840 mm). Beams 2, 3 and 5

8 cracked immediately at release, the other beams did not crack. Crack locations and

9 dimensions are summarized in Fig. 11. For beams tested as part of this study: three of them

10 cracked despite efforts to prevent cracking; and four of them did not crack.

11 DISCUSSION OF RESULTS

12 The adequacy of the beam specimens to resist cracking is summarized in Fig. 12.

13 Beam 5 only cracked along the south end; the end with low amounts of reinforcement.

14 Reinforcement minimally changes the cracking strength due to differences in transformed

15 sectional properties. Therefore, it was determined that the applied tensile stress of Beam 5

16 was very close to the cracking strength. The maximum applied top fiber tensile stress was

17 4.5 f ' ci psi (0.37 f ' ci MPa). The cracking strength estimated by the Collins and Mitchell7

18 expression is 4.3 f ' ci psi (0.36 f ' ci MPa) and the FIB8 expression predicts a strength

19 equal to 4.6 f ' ci psi (0.38 f ' ci MPa).

20 ACI 318-05 §18.4.14 limits the extreme fiber tensile stress to 6 f ' ci psi (0.50 f ' ci

21 MPa) when crack control reinforcement is not provided. Beams 2 and 3 cracked extensively

22 at release (Fig. 11) with an applied tensile stress as low as 5.6 and 5.4 f ' ci (0.47 and

23 0.45 f ' ci ) respectively (Fig. 11). Therefore, the ACI limit of 6 f ' ci psi (0.50 f ' ci MPa)

15
1 was not sufficient to prevent cracking. When adequate crack control reinforcement was

2 provided, crack widths at release were less than or equal to 0.004 in. (0.1 mm).

3 The tensile strength of the concrete in the top flange of the 54 in. (1370 mm) deep

4 beams was not accurately represented by 6x6x21 in. (150x150x530 mm) MOR beams. The

5 measured tensile strengths of the MOR beams were 21% to 52% stronger than the applied

6 stress at release; yet the girders cracked (Fig. 12). Three contributing factors to the seemingly

7 low cracking strength of the beams are discussed in further detail.

8 Effect of Depth on the Flexural Strength of Concrete

9 The strain gradient of a 54 in. (1370 mm) deep beam is over four times shallower than

10 the strain gradient of a 6 in. (150 mm) MOR beam (Fig. 13). As the strain gradient becomes

11 shallower, it approaches the uniform strain condition associated with direct tension. It follows

12 that the resulting tensile strength approaches a value of 4 f ' ci psi (0.33 f ' ci MPa).

13 Also, the greater the depth of a specimen the greater the bleed water at the top surface,

14 hence the weaker the top fiber. For a 54 in. (1370 mm) deep beam, the weakened top fiber is

15 relied upon to resist the applied tensile stress. A MOR beam is nine times less deep and it is

16 tested with its side surface, as cast, in tension.

17 Therefore, the strength of concrete in a 54 in. (1370 mm) deep beam will not be

18 represented by a MOR beam due to bleeding water and strain gradient.

19 Plastic Shrinkage Stresses

20 A full scale 54 in. (1370 mm) deep beam will retain more moisture than a 6 in. (150

21 mm) specimen during curing by virtue of its mass. Therefore, compared to a MOR beam, the

22 differential in shrinkage strains will be relatively higher. Plastic shrinkage stresses are added

23 to the flexural tensile stresses resulting in a lower flexural capacity available to resist

24 cracking. Therefore, it can be assumed that the flexural capacity available to resist cracking

16
1 for a 54-in. deep beam would be lower than that of a MOR beam due to differences in the

2 shrinkage strains.

3 Variability in Concrete Tensile Strength

4 Consider the compilation of modulus of rupture data shown in Fig. 3. When the

5 compressive strength is between 4000 and 5000 psi (28 and 34 MPa) – typical release values

6 – data exists with a tensile strength less than the ACI limit (6 f ' ci psi [0.50 f ' ci MPa]).

7 Admittedly, the number of data points that fall below 6 f ' ci psi (0.50 f ' ci MPa) represents

8 a very small percentage of the total population. However, if these data points represent a

9 worst-case scenario and the girders tested in this study meet the conditions of such a scenario;

10 then cracking would occur regardless of contributions from other factors.

11 Adequacy of ACI 318-08

12 The beams tested as part of the current research program are considered to be bridge

13 members and would typically be designed according to the provisions of the American

14 Association of State Highway and Transportation Officials (AASHTO LRFD)36. AASHTO

15 LRFD has a more stringent crack control requirement than ACI 318. AASHTO requires crack

16 control reinforcement if the tensile stress exceeds 3 f ' ci psi (0.25 f ' ci MPa) at any

17 location (similar to ACI 318-632). Therefore, beams designed according to AASHTO will not

18 experience the cracking problem that has been presented.

19 Nonetheless, the beams in the current study did not crack because they were bridge or

20 building members. They cracked as a result of their depth and the nature of their construction.

21 Within the Precast/Prestressed Concrete Institute (PCI) Handbook37, the following standard

22 sizes are given: double tees as deep as 34 in. (860 mm); rectangular beams as deep as 40 in.

23 (1020); L-beams or inverted tees as deep as 60 in. (1520 mm). Crack control reinforcement is

24 equally necessary for these members as it is for bridge members. Therefore, given the factors

25 that contribute to lowering cracking strength, specifically as they apply to large-scale

17
1 prestressed members, it is proposed that a change to relevant provisions of the ACI 318

2 building code be made in order to allow designers to control cracking with greater assurance.

3 DESIGN RECOMMENDATION

4 Based on the recommendations of Collins and Mitchell7, FIB8, and the results of this

5 study, the following provision is recommended as a replacement to ACI 318 §18.4.1 part (c)4:

6 Where computed concrete tensile strength, ft, exceeds 4 f ' ci psi (0.33 f ' ci MPa),

7 additional bonded reinforcement shall be provided in the tensile zone to resist the

8 total tensile force in concrete computed with the assumption of an uncracked section.

9 The preceding recommendation is based on the fact that the flexural-tensile strength

10 of concrete approaches the direct tensile strength as depth increases; regardless of location. In

11 other words, the tensile stress limit is based on the tensile strength of concrete; and the

12 strength is the same in the middle of the beam as it is in the end region. Therefore, there is not

13 a theoretical basis for specifying a different tensile stress limit at the end of a beam or

14 elsewhere.

15 The intent of the proposed design recommendation is not to prevent cracking at

16 release. The intent is to control cracking. The proposed recommendation would both simplify

17 the current provisions and control cracking with greater assurance. Based on the findings of

18 this experimental program, the explanations provided for changing the ACI 318 building

19 code in 19775 (presented earlier in this paper) are no longer valid for the following reasons:

20 1. 6 f ' ci psi (0.50 f ' ci MPa) is not sufficient to prevent cracking at release for

21 full-scale 54 in. (1370 mm) deep prestressed girders.

22 2. When examining the tensile strength of a prestressed beam at release, a 54 in.

23 (1370 mm) deep girder with high strand eccentricity represents a worst-case

24 scenario; yet this scenario exists. If the intention of §18.4.1 is to control cracking

25 at release, then the provision should be based on aforementioned worst-case

18
1 scenarios. If less deep members (e.g. hollow-core slabs) are able to withstand

2 larger tensile stresses; relief may still be realized according to §18.4.3.

3 3. For the fabrication of a 54 in. (1370 mm) deep girder, the additional expense of

4 adding crack control reinforcement is insignificantly small. The cost of repairing

5 cracks that form at release is much more expensive. Therefore, the benefits of

6 providing adequate crack control reinforcement far out-weigh the alternative.

7 4. 4 f ' ci psi (0.33 f ' ci MPa) is a value just below cracking.

8 5. Although elastic shortening, creep, and shrinkage reduce the level of applied

9 stress; these effects happen after the formation of initial cracks. Therefore, a

10 reduction in stress level after release does not remove the need for crack control

11 reinforcement.

12 CONCLUSIONS

13 Based on the results of this experimental investigation, the following conclusions are

14 drawn:

15 1. The ACI 318 §18.4.1 tensile stress limit is not adequate to prevent some 54 in. (1370

16 mm) deep girders from cracking at prestress release. It is recommended that the

17 tensile stress limit be changed from 6 f ' ci to 4 f ' ci psi (0.5 f ' ci to 0.33 f ' ci

18 MPa) at all locations.

19 2. The tensile strength of the extreme fiber of a 54 in. (1370 mm) deep girder was not

20 accurately represented by a 6x6x21 in. (150x150x530 mm) MOR beam. The tensile

21 strengths of the MOR beams were between 21% and 52% higher than the tensile

22 stresses applied to the girders, yet cracking occurred.

23 3. The amount of crack control reinforcement specified per ACI §18.4.1 was sufficient

24 to limit cracks at release to a width less than or equal to 0.004 in. (0.102 mm).

19
1 4. Stresses predicted at the location of prestress transfer using conventional linear-elastic

2 design assumptions are reliable and accurate.

3 ACKNOWLEDGMENTS

4 The authors would like to thank the Texas Department of Transportation for providing

5 financial support for this research program; and David Birrcher and David Mraz for their

6 assistance with the experimental program. The technical contributions of the project director,

7 Jeff Cotham, are gratefully acknowledged. Opinions, findings, conclusions, and

8 recommendations in this paper are those of the authors.

9 NOTATION:

10 φMOR = strain gradient of a modulus of rupture test

11 φTyIV = strain gradient of a prestressed Type IV girder at release

12 σtens = tensile stress

13 σcomp = compressive stress

14 d = distance from extreme compression fiber to bottom of crack

15 f’c = specified compressive strength of concrete

16 f’ci = specified compressive strength of concrete at time of release

17 fpy = specified yield strength of prestressing steel

18 fr = modulus of rupture of concrete

19 ft = tensile strength of concrete

20 ftop = extreme top fiber stress in tension

21 fy = specified yield strength of non-prestressed reinforcement

22 w = crack width

23

20
1 REFERENCES

2 1. ACI-ASCE Joint Committee 323, “Tentative Recommendations for Prestressed

3 Concrete,” ACI Journal, Proceedings, Vol. 54, Jan. 1958, pp. 545-578.

4 2. ACI Committee 318, Building Code Requirements for Reinforced Concrete (ACI 318-63),

5 American Concrete Institute, Detroit, MI, 1963.

6 3. ACI Committee 318, Building Code Requirements for Reinforced Concrete (ACI 318-77),

7 American Concrete Institute, Detroit, MI, 1977.

8 4. ACI Committee 318, Building Code Requirements for Reinforced Concrete (ACI 318-08),

9 American Concrete Institute, Farmington Hills, MI, 2008.

10 5. ACI Committee 318, “Proposed Revisions to: Building Code Requirements for

11 Reinforced Concrete (ACI 318-71),” ACI Journal, Proceedings, V. 74, No. 1, Jan. 1977,

12 pp. 1-21.

13 6. ACI Committee 224, “Cracking of Concrete Members in Direct Tension,” ACI Journal,

14 Proceedings, V. 83, No. 1, Jan-Feb 1986, pp 3-13.

15 7. Collins, M. P., and Mitchell, D., Prestressed Concrete Structures, Response Publications,

16 Toronto and Montreal, Canada, 1997.

17 8. FIB, Structural Concrete, Textbook on Behaviour, Design and Performance, Updated

18 Knowledge of the CEB/FIP Model Code 1990, Volume 1, Féderation Internationale du

19 Béton, Lausanne, Switzerland, July, 1999.

20 9. Price, W. H., “Factors Influencing Concrete Strength,” ACI Journal, Proceedings, V. 22,

21 No. 6, Feb. 1951, pp. 417-432.

22 10. ASTM C78 – 02, Standard Test Method for Flexural Strength of Concrete (Using Simple

23 Beam with Third-Point Loading), American Society for Testing and Materials, West

24 Conshohocken, PA, 2002.

21
1 11. ASTM C496/C496M – 04, Standard Test Method for Splitting Tensile Strength of

2 Cylindrical Concrete Specimens, American Society for Testing and Materials, West

3 Conshohocken, PA, 2004.

4 12. Carasquillo, R. L.; Nilson, A. H.; and Slate, F. O., “Properties of High-Strength Concrete

5 Subjected to Short-Term Loads,” ACI Journal, Proceedings, V. 78, No. 3, May-June

6 1981, pp. 171-178.

7 13. Gonnerman, H. F., and Shuman, E. C., “Compression, Flexural, and Tension Tests of

8 Plain Concrete,” Proceedings, ASTM, V. 28, Part II, 1928, pp. 527-564.

9 14. Grieb, W. E., and Werner, G., “Comparison of Splitting Tensile Strength of Concrete

10 with Flexural and Compressive Strengths,” Public Roads, V. 32, No. 5, Dec. 1962.

11 15. Houk, “Concrete Aggregate and Concrete Properties Investigation, Dworshak Dam and

12 Reservoir,” Design Memorandum, No. 16, U.S. Army Engineer District, Walla Walla,

13 1965, pp. 203-212.

14 16. Hueste, M. B. D.; Chompreda, P.; Trejo, D.; Cline, D. B.; and Keating, P. B.,

15 “Mechanical Properties of High-Strength Concrete for Prestressed Members,” ACI

16 Structural Journal, V. 101, No. 4, July-Aug. 2004, pp. 457-474.

17 17. Kaplan, M. F., “Flexural and Compressive Strength Concretes as Affected by the

18 Properties of Coarse Aggregates,” ACI Journal, Proceedings, V. 55, No. 11, May 1959,

19 pp. 1193-1208.

20 18. Khan, A. A.; Cook, W. D.; and Mitchell, D., “Tensile Strength of Low, Medium, and

21 High-Strength Concrete at Early Ages,” ACI Materials Journal, V. 93, No. 5, Sept.-Oct.

22 1996, pp. 487-493.

23 19. Mirza, S. A.; Hatzinikolas, M.; and MacGregor, J. G., “Statistical Descriptions of

24 Strength of Concrete,” Journal of the Structural Division, ASCE, V. 105, No. ST6, June

25 1979.

22
1 20. Mokhtarzadeh, A., and French, C., “Mechanical Properties of High-Strength Concrete

2 with Consideration for Precast Applications,” ACI Materials Journal, Vol. 97, No. 2,

3 March-April 2000, pp. 136-147.

4 21. Raphael, J. M., “Tensile Strength of Concrete,” ACI Journal, Proceedings, V. 81, No. 2,

5 March-April 1984, pp. 158-165.

6 22. Shah, S. P., and Ahmad, S. H., “Structural Properties of High-Strength Concrete and Its

7 Implications for Precast Prestressed Concrete,” Journal of the Prestressed Concrete

8 Institute, V. 30, No. 6, Nov.-Dec. 1985, pp. 92-119.

9 23. Walker, S., and Bloem, D. L., “Effects of Aggregate Size on Properties of Concrete,” ACI

10 Journal, Proceedings, V. 57, No. 3, Sept. 1960, pp. 283-298.

11 24. Tuchscherer, R. G., and Bayrak, O., An Investigation of the Tensile Strength of

12 Prestressed AASHTO Type IV Girders at Release, Research Report No. 0-5197-2, Center

13 for Transportation Research, The University of Texas at Austin, March 2007.

14 25. Tucker, J., “Statistical Theory of the Effect of Dimensions and of Method of Loading

15 upon the Modulus of Rupture of Beams,” ASTM, Proceedings of the 44th Annual Meeting,

16 Vol. 41, June 1941, pp. 1072-1088.

17 26. Abrams, D. A., “Flexural Strength of Plain Concrete,” ACI, Proceedings, Vol. 18, 1922, p.

18 20.

19 27. Gonnerman, H. F., and Shuman, E. C., “Flexure and Tension Tests of Plain Concrete,”

20 Director’s Report, Portland Cement Association, Nov. 1928, p. 137.

21 28. Reagel, F. V., and Willis, T. F., “The Effect of the Dimensions of Test Specimens on the

22 Flexural Strength of Concrete,” Public Roads, Vol. 12, No. 2, April 1931, p. 37.

23 29. Wright, P. J. F., “The Effect of the Method of Test on the Flexural Strength of Concrete,”

24 Magazine of Concrete Research, Vol. 4, No. 11, Oct. 1952, pp. 67-76.

23
1 30. ASTM A615/A615M – 06a, Standard Specification for Deformed and Plain Carbon-Steel

2 Bars for Concrete Reinforcement, American Society for Testing and Materials, West

3 Conshohocken, PA, 2006.

4 31. ASTM A416/A416M – 06, Standard Specification for Steel Strand, Uncoated Seven-Wire

5 for Prestressed Concrete, American Society for Testing and Materials, West

6 Conshohocken, PA 2006.

7 32. Kehl, R. J., and Carrasquillo, R. L., Investigation of the Use of Match Cure Technology in

8 the Precast Concrete Industry, Research Report No. 1714-2, Center for Transportation

9 Research, The University of Texas at Austin, Aug. 1998.

10 33. ASTM C39/C39M – 05, Standard Test Method for Compressive Strength of Cylindrical

11 Concrete Specimens, American Society for Testing and Materials, West Conshohocken,

12 PA, 2005.

13 34. ASTM C1231/C1231M – 08, Standard Practice for Use of Unbonded Caps in

14 Determination of Compressive Strength of Hardened Concrete Cylinders, American

15 Society for Testing and Materials, West Conshohocken, PA, 2008

16 35. ASTM C469 – 02, Standard Test Method for Static Modulus of Elasticity and Poisson’s

17 Ratio of Concrete in Compression, American Society for Testing and Materials, West

18 Conshohocken, PA, 2002.

19 36. AASHTO, LRFD Bridge Design Specifications, 4th Edition, American Association of

20 State Highway and Transportation Officials, Washington, D.C., 2007.

21 37. PCI Handbook, Precast and Prestressed Concrete, 6th Edition, Precast/Prestressed

22 Concrete Institute, Chicago, Illinois, 2004.

23

24
1 List of Tables

2 Table 1 – Concrete Mixture Design

3 Table 2 – Summary of Test Specimens

4 List of Figures:

5 Fig. 1 – Type IV Girder that Cracked at Release (courtesy of Jeff Cotham, TxDOT).

6 Fig. 2 – Testing Methods used to Determine Tensile Strength of Concrete.

7 Fig. 3 – Relationship between Modulus of Rupture and Compressive Strength of Concrete

8 (1330 Data Points).

9 Fig. 4 – Reinforcement Configuration and Dimensions of a Typical Type IV Beam Specimen.

10 Fig. 5 – Prestressing Strand Configuration

11 Fig. 6 – Reinforcing Steel Strain Gauge Locations

12 Fig. 7 – Time vs. Temperature Curve at Different Thermocouple Locations (Beam 6).

13 Fig. 8 – Thermocouple Locations used to Match-Cure Control Cylinders

14 Fig. 9 – 54 in. (1370 mm) Tall Cylinder used to Determine MOE in Top Flange.

15 Fig. 10 – Experimental Stress Profile (Beam 7)

16 Fig. 11 – Summary of Crack Spacing and Widths at Prestress Release.

17 Fig. 12 – Measured Strength vs. Cracking Strength of Beam Specimens

18 Fig. 13 – Difference in Strain Gradient between a 6x6 in. (150x150 mm) Beam and Type IV

19 Girder.

25
1 Table 1 – Concrete Mixture Design

Coarse Aggregate
1840 lb/yd3 (1100 kg/m3)
(Crushed Limestone)
Fine Aggregate 1430 lb/yd3 (848 kg/m3)
Type III Cement 611 lb/yd3 (362 kg/m3)
Water 214 lb/yd3 (127 N/m3)
Water/Cement Ratio 0.35
HRWR Admixture
8 oz./Cwt. (0.51 kg/kN)
(Sika Viscocrete 2100)
Retarder (Sika Plastiment) 4 oz./Cwt. (0.26 kg/kN)
2

3 Table 2 – Summary of Test Specimens

Release Prestress Eccentricity f’ci ftop†


Specimens ftop†/ f 'ci Result
Time Force* [kip] [in.] [psi] [psi]
Beam 1 38 days 312 21.6 8100 359 4.0 -
Beam 2 10 hrs. 364 21.6 4500 408 6.1 Cracked
Beam 3 8 hrs. 374 21.6 5000 427 6.0 Cracked
Beam 4 17 hrs. 434 15.1 8200 197 2.2 -
Beam 5 3 days 378 21.6 9500 439 4.5 Cracked
Beam 6 10 hrs. 227 21.6 5400 262 3.6 -
Beam 7 5 days 303 21.6 10,700 351 3.4 -
4 Notes:
5 1 in. = 25.4 mm; 1 kip = 4.448 kN; 1 psi = 6.895 kPa
6 * Prestress force before elastic shortening and shrinkage losses

7 ftop calculated at theoretical transfer length location; therefore, it is greater than the stress
8 calculated at the actual crack location (Fig. 11)

~ 46”

9
10 Fig. 1 – Type IV Girder that Cracked at Release (courtesy of Jeff Cotham, TxDOT).

26
1

P
P/2 P/2

σtens σcomp
σcomp
σtens
σtens

P
P/2 P/2
P
(i) Direct Tension (ii) Split Cylinder (iii) Modulus of Rupture
ft = 4√f’c psi ft = 6√f’c psi ft = 7.5√f’c psi

2 (0.33√f’c MPa) (0.50√f’c MPa) (0.62√f’c MPa)

3 Fig. 2 – Testing Methods used to Determine Tensile Strength of Concrete.

27
Compressive Strength, f’c (MPa)
14 28 41 55 69 83 97 110

Current Study (6x6x21 in. beams)


(150x150x530 mm) 11.0

9.7

Modulus of Rupture, fr (MPa)


8.3

6.9
7.5√f’c
5.5
6√f’c
5√f’c 4.1
4√f’c
2.8
3√f’c
1.4

2 Fig. 3 – Relationship between Modulus of Rupture and Compressive Strength of

3 Concrete (1330 Data Points).

28
1

3
4

5 Fig. 4 – Reinforcement Configuration and Dimensions of a Typical Type IV Beam

6 Specimen.

29
2” (50 mm)

21.6” (550 mm) 21.6” (550 mm)

2” (50mm) 2” (50 mm)


Beams 1, 2, 3, 5, 6, 7 Beam 4
1

2 Fig. 5 – Prestressing Strand Configuration

Instrumented Bar, Typ. Strain Gauge CL

2”
(50 mm)

54”
(1370 mm)

6” 5 Spa @ 1’-0” = 5’-0” Symmetric about centerline


(150 mm)
(5 Spa @ 300 mm = 1.500 m)
3

4 Fig. 6 – Reinforcing Steel Strain Gauge Locations

30
66
150
4 60
140 5 6

130 54
2 3
Temperature (oF)

120 49

Temperature (ºC)
7 1 2
110 1 43
3
8
100 38

90 4 32
5
80 27

70 7 21
6
8
60 16
1 hr 6 hr 11 hr 16 hr 21 hr
Hours after Batch
1

2 Fig. 7 – Time vs. Temperature Curve at Different Thermocouple Locations (Beam 6).

Match-Cure Location
Beams 1, 2, 3, 4, and 7

Match-Cure Location
Beams 5 and 6

4 Fig. 8 – Thermocouple Locations used to Match-Cure Control Cylinders.

31
1

2 Fig. 9 – 54 in. (1370 mm) Tall Cylinder used to Determine MOE in Top Flange.

32
1 ft. from end 2 ft. from end

50 50

40 40

Experimentally
30 30
Measured Data Point

Height, in.
Height, in.

20 20
Best Fit Line
Theoretical Line
10 10

0 0
-1500 -1000 -500 0 500 1000 -1500 -1000 -500 0 500 1000
Stress, psi Stress, psi
1
3 ft. from end 4 ft. from end 5 ft. from end

50 50 50

40 40 40

30 30 30
Height, in.
Height, in.

Height, in.
20 20 20

10 10 10

0 0 0
-1500 -1000 -500 0 500 1000 -1500 -1000 -500 0 500 1000 -1500 -1000 -500 0 500 1000
Stress, psi Stress, psi Stress, psi
Notes:
1 in. = 25.4 mm; 1 kip = 4.448 kN; 1 psi = 6.895 kPa
2

3 Fig. 10 – Experimental Stress Profile (Beam 7)

33
1

Sufficient Insufficient
Crack Control Reinforcement Crack Control Reinforcement
As’ = 2.4 in2 As’ = 0.6 in2

6’-0” 4’-0”
4’-0”
Beam 2
Release Time = 10 hr
f’ci = 4500 psi
ftop = 6.1√f’ci psi
ft = 5.8√f’ci psi 5.6√f’ci psi 5.8√f’ci psi
w = 0.004” w = 0.004” w = 0.013”
d = 14” d = 14” d = 30”

11’-0”
4’-0”
4’-0”
Beam 3
Release Time = 8 hr
f’ci = 5000 psi
ftop = 6.0√f’ci psi ft = 5.7√f’ci psi 5.4√f’ci psi 5.7√f’ci psi
w = 0.003” w = 0.002” w = 0.009”
d = 14” d = 14” d = 25”

4’-0”
Beam 5
Release Time = 3 day
f’ci = 9500 psi
ftop = 4.5√f’ci psi ft = 4.3√f’ci psi
w = 0.005”
d = 16”

Notes:
1 in. = 25.4 mm; 1 in2 = 645 mm2; 1 kip = 4.448 kN; 1 psi = 6.895 kPa
2

3 Fig. 11 – Summary of Crack Spacing and Widths at Prestress Release.

34
1

12 Cracked Cracked
9.6 9.3 9.3 9.5
10
f’top
√f’ci

7.6 7.6 7.5


Tensile Stress Factor,

8
6.1 6.0
ACI Limit
6
4.5
4.0
4 3.6 3.4
2.2
2

0
Beam 1 Beam 2 Beam 3 Beam 4 Beam 5 Beam 6 Beam 7
Max Applied
Applied Extreme
Extreme Fiber Fiber Tensile
Tensile StressStress
Measured
MeasuredConcrete
Concrete Strength
Strength (MOR Specimen)
2

3 Fig. 12 – Measured Strength vs. Cracking Strength of Beam Specimens

4
σtens

φMOR ≈ 13 φTyIV
3
13”
φMOR ≈ 4.3 · φTyIV (330 mm)

φTyIV

54”
(1370 mm)

σcomp
φMOR 6”
(150 mm)
3”
(76 mm)
σtens

5 σcomp

6 Fig. 13 – Difference in Strain Gradient between a 6x6 in. (150x150 mm) Beam and Type

7 IV Girder.

35

View publication stats

You might also like