Sie sind auf Seite 1von 25

Composites: Part B 38 (2007) 119143 www.elsevier.

com/locate/compositesb

Near-surface mounted FRP reinforcement: An emerging technique for strengthening structures


L. De Lorenzis
b

a,*

, J.G. Teng

a Department of Innovation Engineering, University of Lecce, via per Monteroni, 73100 Lecce, Italy Department of Civil and Structural Engineering, The Hong Kong Polytechnic University, Hong Kong, China

Received 27 January 2006; accepted 17 August 2006 Available online 18 October 2006

Abstract Near-surface mounted (NSM) ber-reinforced polymer (FRP) reinforcement is one of the latest and most promising strengthening techniques for reinforced concrete (RC) structures. Research on this topic started only a few years ago but has by now attracted worldwide attention. Issues raised by the use of NSM FRP reinforcement include the optimization of construction details, models for the bond behaviour between NSM FRP and concrete, reliable design methods for exural and shear strengthening, and the maximization of the advantages of this technique. This paper provides a critical review of existing research in this area, identies gaps of knowledge, and outlines directions for further research. 2006 Elsevier Ltd. All rights reserved.
Keywords: A. Polymer matrix composites (PMCs); B. Debonding; C. Analytical modeling; D. Mechanical testing; Near-surface mounted reinforcement

1. Introduction Over the past decade, extensive research has been conducted on the strengthening of reinforced concrete (RC) structures using externally bonded ber-reinforced polymer (FRP) laminates; the technology has also been implemented in a large number of practical projects worldwide. More recently, near-surface mounted (NSM) FRP reinforcement has attracted an increasing amount of research as well as practical application. In the NSM method, grooves are rst cut into the concrete cover of an RC element and the FRP reinforcement is bonded therein with an appropriate groove ller (typically epoxy paste or cement grout). What is herein called NSM reinforcement was previously given other names such as grouted reinforcement [1], or embedded reinforcement [2,3].

Corresponding author. Tel.: +39 0832 297241; fax: +39 0832 297279. E-mail address: laura.delorenzis@unile.it (L. De Lorenzis).

Examples of the use of NSM steel rebars in Europe for the strengthening of RC structures date back to the early 1950s [1]. More recent applications of NSM stainless steel bars for the strengthening of masonry buildings and arch bridges have also been documented (e.g. [4]). The advantages of FRP versus steel as NSM reinforcement are better resistance to corrosion, increased ease and speed of installation due to its lightweight, and a reduced groove size due to the higher tensile strength and better corrosion resistance of FRP. Compared to externally bonded FRP reinforcement, the NSM system has a number of advantages: (a) the amount of site installation work may be reduced, as surface preparation other than grooving is no longer required (e.g., plaster removal is not necessary; irregularities of the concrete surface can be more easily accommodated; removal of the weak laitance layer on the concrete surface is no longer needed); (b) NSM reinforcement is less prone to debonding from the concrete substrate; (c) NSM bars can be more easily anchored into adjacent members to prevent debonding failures; this feature is particularly attractive in the exural strengthening of beams and columns in rigidly-jointed

1359-8368/$ - see front matter 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.compositesb.2006.08.003

120

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

frames, where the maximum moments typically occur at the ends of the member; (d) NSM reinforcement can be more easily pre-stressed; (e) NSM bars are protected by the concrete cover and so are less exposed to accidental impact and mechanical damage, re, and vandalism; this aspect makes this technology particularly suitable for the strengthening of negative moment regions of beams/ slabs; (f) the aesthetic of the strengthened structure is virtually unchanged. Due to the above advantages, the NSM FRP method is in many cases superior to the externally bonded FRP method or can be used in combination with it, provided that the cover of the member is sufciently thick for grooves of a desirable size to be accommodated. The existing knowledge on the NSM FRP method is much more limited than that on the externally bonded FRP method, as reected by the absence of relevant provisions in the existing guidelines on the FRP strengthening of concrete structures issued by b [5] and ACI-440 [6]. However, the international engineering community has become increasingly aware of the practical advantages of this method, which has led to accelerations of research and practical applications worldwide. Both ACI-440 and b are currently considering revisions to their documents to include NSM-related provisions. Against this background, this paper provides a critical review of existing research in this area, identies gaps of knowledge, and outlines directions for further research. This paper focuses on research work on the structural aspects of NSM strengthening of concrete structures. Discussions on some signicant practical applications of the NSM method, on NSM strengthening of masonry and timber structures, and on durability-related aspects are given in Ref. [7]. Space limitation does not allow them to be discussed in this paper. 2. Materials and systems 2.1. FRP reinforcement In most existing studies, carbon FRP (CFRP) NSM reinforcement has been used to strengthen concrete structures. Glass FRP (GFRP) has been used in most applications of the NSM method to masonry and timber structures. The present authors are not aware of any study or practical application in which aramid FRP (AFRP) was used. The tensile strength and elastic modulus of CFRP are much higher than those of GFRP, so for the same tensile capacity, a CFRP bar has a smaller cross-sectional area than a GFRP bar and requires a smaller groove. This in turn leads to easier installation, with less risks of interfering with the internal steel reinforcement, and with savings in the groove-lling material. FRP bars can be manufactured in a virtually endless variety of shapes. Hence, the NSM FRP reinforcement may be round, square, rectangular and oval bars, as well as strips (Figs. 1 and 2). For brevity, the term bars is

Fig. 1. Types of FRP bars for NSM applications.

Fig. 2. Dierent NSM systems and nomenclature.

used herein as a generic term encompassing all crosssectional shapes, while the term strips is reserved for thin narrow strips. Dierent cross-sectional shapes have dierent advantages, and oer dierent choices for practical applications. For example, square bars maximize the bar sectional area for a given size of square groove while round bars are more readily available and can be more easily anchored in pre-stressing operations. Narrow strips

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

121

maximize the surface area-to-sectional area ratio for a given volume and thus minimize the risk of debonding, but require a thicker cover for a given cross-sectional area. In practical applications, the choice depends strongly on the constraints of a specic situation, such as the depth of the cover, and the availability and cost of a particular type of FRP bar. FRP bars are also manufactured with a variety of surface textures, which strongly aect their bond behaviour as NSM reinforcement. Their surface can be smooth, sand-blasted, sand-coated, or roughened with a peel-ply surface treatment. Round bars can also be spirally wound with a ber tow, or ribbed [8]. 2.2. Groove ller The groove ller is the medium for the transfer of stresses between the FRP bar and the concrete. In terms of structural behaviour, its most relevant mechanical properties are the tensile and shear strengths. The tensile strength is especially important when the embedded bars have a deformed surface, which produces high circumferential tensile stresses in the cover formed by the groove ller (simply referred to as the cover or the epoxy cover hereafter) as a result of the bond action. In addition, the shear strength is important when the bond capacity of the NSM reinforcement is controlled by cohesive shear failure of the groove ller. The eect of the modulus of elasticity of the groove ller has never been experimentally investigated. The most common and best performing groove ller is a two-component epoxy. Low-viscosity epoxy can be selected for strengthening in negative moment regions as the epoxy can be poured into the grooves. For other cases, a high-viscosity epoxy is needed to avoid dripping or owing-away. The addition of sand to epoxy can increase the volume, control the viscosity, lower the coecient of thermal expansion, and raise the glass transition temperature. A drawback of this addition seems to be reduced adhesion at the barepoxy interface for a smooth bar surface [2]. The use of cement paste or mortar in place of epoxy as a groove ller has recently been explored in an attempt to lower the material cost, reduce the hazard to workers, minimize the environmental impact, allow eective bonding to wet substrates, and achieve better resistance to high temperatures and improved thermal compatibility with the concrete substrate. However, cement mortar has inferior mechanical properties and durability, with a tensile strength an order of magnitude smaller than that of common epoxies. Results of bond tests and exural tests [9,10] have identied some signicant limitations of cement mortar as a groove ller. Given these limitations and the very limited data available, the rest of this paper is focused on epoxy-bonded NSM FRP reinforcement only, except when future research needs are discussed. Nevertheless, tests on NSM FRP reinforcement using cement grout as

the groove ller, if available, are included in Tables 14 for completeness. 2.3. Groove dimensions Fig. 2 shows several congurations of NSM FRP reinforcement, where db is the nominal diameter of a round bar, and tf and hf are the thickness/width and the height of an FRP strip or rectangular bar respectively. The groove width bg, the groove depth hg, the net distance between two adjacent grooves ag, and the net distance between a groove and the beam edge ae are all relevant construction parameters, which can inuence the bond performance and hence the structural behaviour. For round bars, De Lorenzis [11], based on results of bond tests with square grooves (bg = hg) and dening k = bg/db, proposed a minimum value of 1.5 for k for smooth or lightly sand-blasted bars and a minimum value of 2.0 for k for deformed bars. Parretti and Nanni [12] suggested that both bg and hg should be no less than 1.5db. For NSM strips, Blaschko [13] suggested that the depth and width of the cut groove should be about 3 mm larger than the height and thickness of the corresponding FRP strip respectively, so to obtain an adhesive layer thickness of about 12 mm. Also for NSM strips, Parretti and Nanni [12] recommended that the minimum width of a groove be no less than 3tf and the minimum depth be no less than 1.5hf. For a more detailed discussion, see the section on bond behaviour. In the existing studies, NSM strips were bonded using epoxy either along all four sides of the strip surface [14,15], or along three sides of the strip surface only [16,17] (Fig. 2). Due to the large width to thickness ratio of the strips, the reduction in the bond surface in the latter case is negligible. In existing tests on NSM square bars [18], only three sides of the bar surface were bonded to the concrete member. 2.4. Groove position If a single NSM bar is to be provided to the tension side of an RC member, it should naturally be centrally located over the beam width. When two or more NSM bars need to be provided, then the distance between two adjacent NSM bars and the distance between the edge of the member and the adjacent bar become important design parameters. The eect of these parameters is discussed in the section on bond behaviour. 2.5. Constructional aspects Compared with the use of externally bonded FRP laminates, the need to cut grooves into the concrete member in the construction process of the NSM method is the key difference. A detailed discussion of the construction process can be found in Ref. [7].

122

Table 1 Summary of existing experimental work on the bond behaviour of NSM FRP reinforcement bonded to concrete
Round bars [2] Test method Specimen shape and dimensions Direct pull-out Epoxy-lled plastic pipe, diameter = 51 mm, one bar embedded concentrically Epoxy or epoxy + sand 1520 from manufacturer (value with no sand) Concrete block, 152 152 203 mm, two NSM bars on opposite sides 34 Epoxy 13.8 from manufacturer Two concrete blocks, each 150 150 300 mm, two NSM bars on opposite sides [38] [3] [21] Beam pull-out Inverted T beam, total height = 254 mm, ange width = 254 mm, net span = 1.07 m, one NSM bar 28 Epoxy 13.8, from manufacturer [23] Direct pull-out C-shaped specimen, external dimensions are 300 300 400 mm, one NSM bar 22 Epoxy 27.4, from testing C-shaped specimen, external dimensions are 300 300 400 mm, one NSM bar 22 Epoxy or cement paste 27.4, from testing (epoxy) 6.3, bending tensile strength from testing (cement) CFRP/round/ SWS GFRP/round/ RB [26] Square bars [9] Beam pull-out Beam with two halves connected by a steel hinge, net span = 1.73 m, cross-sectional dimensions N/Aa, one NSM bar N/Aa Epoxy or cement paste 31, from manufacturer (epoxy)e 9, bending tensile strength from manufacturer (cement) Strips [13,16] Direct pull-out Concrete block, 300 300 1100 mm, one NSM strip per test Concrete block, 150 150 350 mm, one NSM strip per test [20] [14] Beam pull-out RILEM beam pullout test

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

fc0 (MPa) Groove-ller Direct tensile strength of groove ller (MPa)

20 Epoxy N/Aa

32 and 46 Epoxy 33.3 from testing

29 cube strength Epoxy 42.6 from testing

35, 45, 70 Epoxy 1622, from testing

Type of FRP/ crosssectional shape/ surface conguration Nominal db or tf hf of FRP (mm) bg hg (mm) Groove surface Test variables

CFRP/ round/SM CFRP/ round/SB

CFRP/ round/SB

CFRP/round/ SB

9.5

11

9.5

CFRP/round/ SB CFRP/round/ RB GFRP/round/ RB 9.5 and 12.7

CFRP/round/ SWS GFRP/round/ RB

CFRP/square/SM CFRP/square/SC

CFRP/strip/ roughened

CFRP/strip/ roughened

CFRP/strip/ N/Aa

7.5 and 9.5

7.5 and 9.5

10 10

(1.22) 20

5 16

1.4 9.3

Bar surface, type of epoxy BE-I

14.3 19.0 Saw-cut BL

16 25 Saw-cut BL

bg = hg, k = 1.122.67 Saw-cut BL, type of bar, bar diameter, k BE-I, SP-E, SP-C1, SP-C2

bg = hg, k = 1.242.50 Pre-formed BL, type of bar, k EC-I, SP-C1

Observed bond failure modesb

Concrete shear at edgec, BE-I

Concrete shear failurec

b g = h g, k = 1.242.50 Saw-cut BL, type of bar, grooveller SP-C1, BE-C

16 16 Saw-cut BL, bar surface conguration, groove-ller N/Aa (indicated as concrete failure and delamination)

(2.84.3) (2123) average 3.3 22 Saw-cut BL, type of strip, fc0 , type of loading, ae BE-C, SP-ED

9 22 Saw-cut BL

3.3 15 Saw-cut BL, concrete strength BEd

BE-I

Acronyms: BL = bond length; RB = ribbed; SB = sand-blasted; SC = sand-coated; SM = smooth; SWS = spirally wound and sand-coated. a N/A = not available. b For acronyms see Fig. 3. c Failure mode inuenced by specimen details. d Unspecied whether BE-I or BE-C. e Tensile test method unspecied.

Table 2 Local bond strength of NSM systems (saw-cut grooves) Type of bar (material, cross-sectional shape) CFRP, round CFRP, round Surface conguration Sand-blasted Ribbed Nominal db or tf hf (mm) 9.5 12.7 9.5 k or bg hg (mm) 1.342.67 1.25 1.33 1.59 2.12 1.36 1.64 2.18 1.50 2.00 2.50 N/Aa 3.3 15 Average 3.3 22 9 22 1.59 2.21 1.64 2.27 1.50 N/Aa Groove ller Epoxy Direct tensile strength of groove ller (MPa) 13.8 From manufacturer 27.4 From testing fc0 (MPa) 28 22 Local bond strength (MPa) 8.6 9.7 11.2 15.4 16.6 9.1 10.0 12.5 18.0 20.8 21.9 9.0c 19.8 20.0d 1012 9.7 6.4 8.0 8.3 6.7 4.3c Mainb failure mode BE-I SP-C1 Reference [21] L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143 [26]

GFRP, round

Ribbed

9.5

CFRP, round

Spirally wound and sand-coated Smooth N/Aa Roughened Roughened

7.5

CFRP, square CFRP, strip

10 10 1.5 9.6 (1.22) 20 5 16 9.5 9.5 7.5 10 10

31 From manufacturer 1622 From testing 33.3 From testing 42.6 From testing Cement paste 6.3, Bending tensile strength from testing

N/Aa 35, 45, 70 32 and 46 29 Cube strength 22

N/Aa BE BE-C BE-I SP-C1, BE-C

[9] [14] [16] [20] [26]

CFRP, round GFRP, round CFRP, round CFRP, square


a b c d

Ribbed Ribbed Spirally wound and sand-coated Sand-coated

9, Bending tensile strength from manufacturer

N/Aa

N/Aa

[9]

N/A = not available. In some cases more than one failure mode were concurrent. Taken from the specimen with the shortest bonded length (100 mm). With no inuence of edge distance (i.e. a0e P 150 mm).

123

124

Table 3 Summary of existing experimental work on exural strengthening with NSM FRP reinforcement beams with limited bonded lengths, slabs, and columns Beams with limited bonded lengths Reference Test method [17] 3-Point bending [19] [20] 4-Point bending Slabs [2] Patch loading at centre [42] 3-Point bending [43] Uniform line load at mid-span or at 330 from free end of cantilever slab 1180 406 PC Columns [42] Horizontal load at top

Cross-sectional shape and dimensions (mm) Net span (m) Shear span (m) fc0 (MPa) Groove ller Direct tensile strength of groove ller (MPa) Type of FRP/crosssectional shape/ surface conguration FRP nominal db or tf hf (mm) Number of FRP bars

T, total height = 300, web height = 250, ange width = 300, web width = 150 2.5 48 Epoxy N/Aa

Rectangular, 150 300

5400 229 RC, simply supported

1800 229 PC, simply supportedc

7600 460

600 600

3 28 N/Aa

3 41 N/Aa

7.9 56 N/Aa

35 42.6 from testing

4.9 (mid-span), 1.8 (cantilever) 4550 N/Aa

Height 1.83.4 56 N/Aa

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

CFRP/strip/ N/Aa

CFRP/ round/RB

CFRP/ strip/PP

CFRP/round/SM

CFRP/round/SM CFRP/strip/N/Aa

CFRP/round/SB

CFRP/round/N/Aa CFRP/strip/N/Aa CFRP/round/RB 10; 1.4 25; 9.5 10; 18; (mid-span) 6; 6; 8 (cantilever)

CFRP/round/ SB

1.2 25 1

9.5 1

4 16 1

3 4 Bars every 102 mm o.c. (long.), 1 bar every 102 mm o.c. (transv.) 172 1550 9.5 13 16-mm bars @102 mm o.c. (long.) and 204 mm o.c. (transv.)

9.5, 1.2 25 1 Bar every 102 mm o.c. (long.), 1 strip every 51 mm o.c. (long.) 172 (bars), 130 (strips) 1550 (bars), 1790 (strips) 16 25, 3.2 29 None

11 20 bars (one every 381 mm o.c.)

11 3 on each face or 7 on each face 119 1240 14 19 Four 19-mm bars

Elastic modulus of FRP (GPa) Tensile strength of FRP (MPa) bg hg (mm) Steel tension reinforcement

150 2000 5 25 Two 10-mm bars

111 1918 18 30

151 2068 8 22 Two 12-mm bars

119 1240 14 19 25-mm bars @127 mm o.c. (long.) and 13-mm bars @457 mm o.c. (transv.) Bar length 6 m

147; 150; 111 1970; 2000; 1918 16 30; 5 25; 16 30 Five 16-mm mild bars (mid-span), four 16-mm mild bars (cantilever) + twelve 15-mm 7-wire pre-stressing strands N/Aa

Bar anchorage lengthb (mm) Test variables Observed failure modes Increase in ultimate load (%)

1501200

1501200

01150

Bar anchorage length Debonding by EC-Cb at cut-o and MMR, BR 054

Concrete Splittingc (MMR) 041

CCS (cut-o), CCS (MMR) 0106

Termination: 152 from support, 762 from free edge Punching shear + DBd 15

N/Aa

381 mm in the footing Number of bars BR, CC

CC + DBd (bars), CC (strips) 300

BR

Type of bar CC

27

36; 43; 39

102, 177

Acronyms: BR = bar rupture, CC = concrete crushing, CCS = concrete cover separation; DB = debonding; MMR = maximum moment region; RC = reinforced concrete; PC = pre-stressed concrete; PP = roughened with peel-ply surface treatment; RB = ribbed; SB = sand-blasted; SM = smooth. a N/A = not available. b See Fig. 3. c Not clear whether it was CCS. d Debonding mechanism not clear.

Table 4 Summary of existing experimental work on exural strengthening with NSM FRP reinforcement beams
Reference [61] Test method Cross-sectional shape and dimensions (mm) Net span (m) Shear span (m) fc0 (MPa) Groove ller Direct tensile strength of groove ller (MPa) Type of FRP/ cross-sectional shape/surface conguration FRP nominal db or tf hf (mm) Number of FRP bars Elastic modulus of FRP (GPa) Tensile strength of FRP (MPa) bg hg (mm) Steel tension reinforcement ratio (%) Cut-o distance from support (mm) Test variables 4-Point bending Rectangular, 457 152, 914 356 2.4, 7.5 0.8, 2.5 N/Aa Epoxy 30 T, total height = 406, web height = 305, ange width = 381, web width = 152 3.9 1.83 36 Epoxy 13.8 Rectangular, 200 400 Rectangular, 200 300 Rectangular, 200 500, 600 500 2.8, 7.5 1.15, 3.25 44 Epoxy 33.3 [44] [11] [10] [16] [41] 3-Point bending T, total height = 300, web height = 250, ange width=300, web width=150 2.5 48 Epoxy 48 (bars) 70 (strips) [62] 4-Point bending Rectangular, 152 304 188 Rectangular, 200 120 Rectangular, 100 170 180 [63] [15]

4 1.75 15 Epoxy 27.4

3.6 1.3 61 Epoxy and cement grout N/Aa

1.3 0.5 37 Epoxy N/Aa

1.1 0.5 2063 Epoxy N/Aa

1.5 0.5 46 Epoxy

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

1622 from testing

CFRP/round/ SM

GFRP/round/RB CFRP/round/SB

CFRP/round/ SWS

4.75, 6.35 4, 11 122 1326 bg = 10.2, varying hg 1.14, 1.19

12.7 (GFRP), 9.5 and 12.7 (CFRP) 2 38.6 (GFRP) 773 (GFRP) 19 19 (db = 9.5) 25 25 (db = 12.7) 0.89

7.5 1, 2 175 2214 16 16 0.380.64

CFRP/square/ SM CFRP/square/ SC 10 10 2 230 4140 N/Aa 0.67

CFRP/strip/ roughened

2 20 3, 11 156 1813 3.3 23 0.63, 0.84

CFRP/round/SW CFRP/strip/N/A CFRP/strip/ N/A GFRP/strip/N/A 9.5, 2 16, 1.2 25, 2 20 1, 2 (CFRP), 5 (GFRP) 122.5, 140, 150, 45 1408, 1525, 2000, 1000 18 30, 6.4 19, 6.4 25, 6.4 25 0.48

CFRP/strip/ roughened

CFRP/round/ N/Aa

CFRP/strip/N/A

0.25 15.5 1, 2 136.6 1656 6.4 19 0.831.74

7 (net); 8 (external) 1, 3 201 1940 N/Aa 0.28, 0.57

1.5 9.6 13 159 2740 4 12 0.330.84

N/Aa

Extended over supports Type of FRP bar, bar diameter DBb of NSM SB bar, CCS (MMR)

Extended over supports Steel ratio, number of FRP bars CC, CCS (MMR), edge failure 2161

300 or extended over supports NSM bar length, groove ller DBc of NSM bars, BR (when extended over supports) 5692

150, 300

50

Extended over supports Section width, steel ratio, number of FRP bars CC, BR

50

50

Beam size, steel ratio, groove size CC, secondary debonding and partial BR 2050

End anchorage, type of loading CCS from cut-o, BR

Type of FRP bar

Observed failure modes

Increase in capacity (%)

2644

6782

CCS (MMR) (CFRP round bars and GFRP strips), BR (CFRP strips) 6999

Number of FRP bars, concrete strength CC, DB (described as rod slippage) 140430

Steel ratio, number of FRP bars CCSd

1555

7898

Acronyms: BR = bar rupture, CC = concrete crushing, CCS = concrete cover separation; DB = debonding; MMR = maximum moment region; RB = ribbed; SB = sand-blasted; SC = sand-coated; SM = smooth; SW = spirally wound; SWS = spirally wound and sand-coated. a N/A = not available. b Debonding mechanism BE-I (Fig. 3), see Fig. 7a. c Debonding mechanism not clear.

125

126

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

3. Bond behaviour 3.1. Summary of existing work The bond between an NSM bar and the substrate material plays a key role in ensuring the eectiveness of the NSM strengthening method. The performance of the bond depends on a number of parameters: the groove and the bar dimensions, the tensile and shear strengths of the concrete and the groove ller, the bar cross-sectional shape and surface conguration, and the degree of roughness of the groove surface. This large number of parameters calls for extensive laboratory characterization, as well as analytical and numerical modelling. Table 1 provides a summary of the existing experimental data on bond behaviour. Note that the tests in some existing studies [17,19,20], although addressing bond-related aspects, were conducted on beams under bending and are therefore discussed in a later section. 3.2. Failure modes and mechanisms The bond tests summarized in Table 1 identied dierent possible bond failure modes of NSM systems (Fig. 3). These modes are described in some detail below and the underlying mechanisms examined. 3.2.1. Bond failure at the barepoxy interface This mode may occur as either pure interfacial failure (BE-I), or as cohesive shear failure in the groove ller (BE-C). The BE-I mode is critical for bars with a smooth or lightly sand-blasted surface, i.e. when the degree of surface deformation is insucient to provide mechanical interlocking between the bar and the groove ller and the bond resistance relies primarily on adhesion between the bar and the ller. For round bars, this mode becomes critical if the groove size is suciently large to avoid splitting failure of the groove ller. For epoxy and concrete of moderate strengths, De Lorenzis and Nanni [21] estimated that for lightly sand-blasted round bars, a k value of 1.5 was enough to prevent splitting failure of the epoxy cover. For round bars, cracking of the epoxy cover (Fig. 3) produced by the radial components of the bond stresses can accelerate the occurrence of a BE-I failure. The BE-C failure mode was observed for NSM strips with a roughened surface [13,16]. This mode occurs when the shear strength of the epoxy is exceeded. Inter-laminar shear failure within the bar, although theoretically possible, has never been observed. Shearing-o of ribs in ribbed bars has never been reported as a failure mode itself, unlike in the case of FRP ribbed bars as internal reinforcement for concrete [22]. However, in some tests [21], the surface of the ribbed bars was found to have been damaged after bond failure, indicating that this could be an upper-bound failure mode.

3.2.2. Bond failure at the epoxyconcrete interface Bond failure at the epoxyconcrete interface may occur as pure interfacial failure (EC-I), or as cohesive shear failure in the concrete (EC-C). The EC-I failure mode was found to be critical for pre-cast grooves [23]. For spirally wound bars or ribbed bars with low rib protrusions, this was found to be the critical failure mode whenever the groove was preformed, independent of the value of k. For ribbed bars with high rib protrusions, this mode was found to be critical only for k values larger than a minimum value (equal to approximately 2.00 for ribbed bars in epoxy), and for lower k values, splitting failure of the epoxy cover dominated. The EC-C failure mode has never been observed in bond tests, but it has been observed in bending tests on beams (Fig. 3, see also Fig. 7f) within the strengthened region [11,20] or at the bar cut-o point [17]. The latter authors considered this failure mode in their theoretical model for debonding of NSM strips. 3.2.3. Splitting of the epoxy cover Longitudinal cracking of the groove ller and/or fracture of the surrounding concrete along inclined planes is herein referred to as cover splitting. This was observed to be the critical failure mode for deformed (i.e. ribbed and spirally wound) round bars. The mechanics of cover splitting bond failure in an NSM system is similar to that of splitting bond failure of steel deformed bars in concrete, on which a good understanding has been developed from decades of research [24]. For an NSM FRP bar, the radial component of the bond stresses is balanced by circumferential tensile stresses in the epoxy cover which may lead to the formation of longitudinal splitting cracks of the cover. The concrete surrounding the groove is also subjected to tensile stresses and may eventually fail when its tensile strength is reached, causing fracture along inclined planes. Whether fracture in the concrete occurs before or after the appearance of splitting cracks in the cover or even after the complete fracture of the cover, depends on the groove size and the tensile strengths of the two materials. The tensile strength of epoxy is one order of magnitude larger than that of concrete, but the epoxy cover thickness for NSM FRP reinforcement is one order of magnitude smaller than the thickness of concrete cover to internal steel reinforcement in an ordinary RC member. Moreover, longitudinal steel reinforcement in RC beams benets from the restraint of shear links, but this restraint is not available for NSM longitudinal reinforcement, unless external restraint of some form (e.g. FRP U jackets as shear reinforcement) is provided. These factors explain why cover splitting is a likely bond failure mode for an NSM system. Figs. 4bc illustrate how the bond mechanism of an NSM system can be modelled in the plane perpendicular to the bar axis, as further explained later. These gures also clarify the dierence in bond mechanism between round bars

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

127

Fig. 3. Bond failure modes of NSM systems observed in bond tests.

and strips. In the latter case, the normal component of the bond stresses is transverse to the thick lateral sides of the groove [13,25] so that splitting failure is less likely to occur. The dierent patterns of cover splitting failure of NSM systems are shown in Fig. 3. When the k ratio is very low (e.g. specimens in Ref. [21] with k = 1.121.18), failure is limited to the epoxy cover and involves little damage in

the surrounding concrete (SP-E failure). For higher values of k, failure involves a combination of longitudinal cracking in the epoxy cover and fracture of the surrounding concrete along inclined planes (SP-C-1 failure); concrete fracture starts as soon as the epoxy cover cracks and the tensile stresses are redistributed [26]. The inclined fracture planes in the concrete have been observed to form an angle (b in Fig. 3) of approximately 30 with the horizontal. For

128

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

Fig. 4. Schematics of the bond behaviour of NSM FRP reinforcement: (a) bond stresses in the longitudinal plane; (b) normal stresses in the transverse plane generated by a round bar; (c) normal stresses in the transverse plane generated by a rectangular bar.

large groove depths and/or when the tensile strength ratio between concrete and epoxy is small, fracture of concrete may occur before the epoxy crack has reached the external surface (SP-C-2 failure). The bond failure modes discussed above are for an NSM bar located centrally in a wide member, where edge eects are unimportant. When an NSM bar is close to the edge of a concrete member, failure involves the splitting of the edge concrete (SP-ED failure) [16]. In tests, this failure mode was found to occur when a0e < 20 mm [16], with the angle b 0 dened in Fig. 3 ranging from 45 to 70. The bond strength associated with the SP-E mode is expected to depend strongly on the tensile strength of the epoxy, whereas those associated with the SP-C-1 and SPC-2 modes are expected to depend strongly on the concrete tensile strength. In all cases, the bond strength is expected to increase with the cover thickness of the NSM bar (i.e. the groove depth). However, the rate of strength increase with the groove depth has been observed to reduce after the groove depth exceeds about 2 times the diameter, around which the SP-C-2 mode replaces the SP-C-1 mode as the critical bond failure mode. The bond strength of the SP-E mode also increases as the surface deformations become less pronounced [26]. 3.3. Eect of groove detailing on bond performance The eect of the groove width-to-depth ratio on the bond performance has not yet been investigated in detail. Through nite element modelling, the debonding load of an NSM strip was found to increase with the groove width

[17]. As the groove depth was kept constant, an increased groove width implied a larger interfacial area between epoxy and concrete. This in turn implied a larger debonding load since debonding was assumed to occur by cohesive shear failure in the concrete at the epoxyconcrete interface [17]. By simplied analytical modelling for deformed round bars [25], the cracking load of the epoxy cover was found to decrease with an increase in the groove width-to-depth ratio (for a given depth) but the failure load was found to remain substantially unchanged due to failure being in the concrete along fracture surfaces nearly independent of the groove width-to-depth ratio. The rst result was conrmed by Hassan and Rizkalla [19] through nite element modelling. It was also found that the tensile stresses in the concrete decrease with an increase in the groove width, which implies a larger concrete cracking load (but not necessarily a larger failure load). However, experimental evidence on the eect of the groove width-to-depth ratio on bond performance is still lacking. From bond tests on NSM strips, Blaschko [13,16] indicated that a minimum a0e a0e ae bg =2 of about 20 mm was required to avoid a splitting failure of the concrete corner, and for a0e values larger than 30 mm, no cracks were observed in the concrete at bond failure. He suggested that a0e be no less than 30 mm or the maximum aggregate size, whichever is greater. The maximum aggregate size was suggested as a limit to avoid damaging the concrete during the cutting of the groove. In his bond tests, a0e did still inuence the bond behaviour until the maximum investigated value of 150 mm, beyond which no further inuence was assumed.

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

129

Based on nite element modelling for round deformed bars, Hassan and Rizkalla [19] suggested a minimum ag and a minimum ae of two and four bar diameters respectively, regardless of the groove width. However, one of the beams tested by De Lorenzis [11], which was strengthened with NSM spirally wound round bars with ag = 30 mm (i.e. about 1.8 times the groove size and 3.6 times the bar diameter) and ae = 69 mm (i.e. about 4.3 times the groove size and 8.6 times the bar diameter), failed by debonding of the NSM bars involving the spalling of the concrete cover of the longitudinal steel reinforcement along the edges. This test thus suggested that the minimum values for ag and ae as specied in Ref. [19] are insucient to eliminate interactions between an NSM bar and the edge of a beam. 3.4. Local bond strength 3.4.1. Experimental results In any type of bond test, the average bond strength usually decreases with increases in the bond length, as a result of the non-uniform distribution of bond stresses (Fig. 4a). The local bond strength refers to the maximum value of bond stress that the interface can resist, in contrast to the overall bond strength (or simply bond strength as used in this paper) which refers to the maximum transferable load of the joint. The local bond strength must then be obtained either from very short specimens or from a long specimen by elaborative strain (and/or slip) measurements. Several authors studied the local bond strengths of NSM systems [13,27,16,20,21,25]. A summary of their values, reported in Table 2, allows the following observations to be made:  the local bond strength of the cover splitting mode (deformed bars), as expected, is higher if the groove depth is larger and the bar surface deformation less pronounced;  the local bond strength of the barepoxy interfacial (BEI) failure mode, which was observed for sand-blasted bars, is not inuenced by the groove size and is lower than that for deformed bars;  the local bond strengths of NSM strips from two test series by dierent authors [14,16] are very close to each other and are comparable to that of spirally wound bars [26]; by contrast, the local bond strength obtained from a third test series [20] is notably lower. Available information on the bond behaviour of square bars is still very limited. Nordin and Taljsten [9] reported average bond strengths from specimens whose lengths were at least 10 times the width of the square cross-section but no local bond strengths were reported. An unusual aspect of the results reported by these authors is that the average bond strength increased with the bond length. De Lorenzis [25] reported the local bond strengths at the epoxyconcrete interface for specimens with dierent types of bars tested by De Lorenzis et al. [23]. As EC-I failure

was the critical mode, the surface conguration of the bar did not have a signicant eect on the bond behaviour and the dierence in local bond strength between ribbed and spirally wound bars was basically due to the dierent diameters, and hence to the dierent groove size corresponding to a given k value. The local bond strength was found to decrease with an increasing groove size. 3.4.2. Theoretical models for NSM strips It is interesting to compare the experimental local bond strengths of NSM strips reported by Sena Cruz and Barros [27] with the predictions by the formula proposed by Blaschko [13], and with those given by Hassan and Rizkallas theoretical model [17]. Blaschkos formula [13] is given by: p smax 0:2 4 a0e saf a0e 6 150 mm 1 where saf is the shear strength of the epoxy. Hassan and Rizkallas formula [17] is given by: smax fc0 fc0 fct fct 2

where fc0 and fct are the (cylinder) compressive and tensile strengths of concrete, respectively. The two formulae relate the local bond strength to dierent parameters, consistent with their own experimental observations: Blaschko observed cohesive shear failure in the epoxy and studied the eect of a0e , whereas Hassan and Rizkalla observed cohesive shear failure in the concrete (hence, their value of smax is the shear strength of concrete). The following dierences between the two formulae should also be noted: (a) Blaschko performed pull-out bond tests to provide the experimental basis, while Hassan and Rizkalla conducted exural tests on RC beams embedded with bars of varying lengths; (b) Blaschkos formula was calibrated with bond test results, while Hassan and Rizkallas formula was derived from Mohrs circle for the pure shear stress state, which, when used in nite element modelling, yielded predictions of the debonding load in good agreement with test results. The 95 percentile characteristic value of saf was indicated by Blaschko [16] to vary between 20 and 25 MPa for common highly lled, two-component epoxies. According to the tests by Blaschko, the ratio between the characteristic and the average values of saf is about 0.89, hence the average value of saf of common epoxies can be assumed to vary between 22.5 and 28.1 MPa. For a0e 150 mm (i.e. with no edge eect), Eq. (1) thus yields a local bond strength ranging between 15.8 and 19.8 MPa, whose upper bound practically coincides with the values in Table 2 from tests in [14,16]. This is as expected as Eq. (1) was calibrated using test results with an average value of 28.7 MPa for saf, very close to the upper bound of the average value range mentioned above. For fc0 p ranging between 20 and 40 MPa and taking fct as 0:53 fc0 [28], Eq. (2) predicts local bond strengths between 2.1 and 3.1 MPa. The large dierence between

130

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

the predictions of Eqs. (1) and (2) is a result of the dierent materials controlling the failure (epoxy for Eq. (1) and concrete for Eq. (2)) and thus the dierent interfaces that these two formulae correspond to; the concrete shear strength is much smaller than that of the adhesive. 3.4.3. Theoretical models for NSM round deformed bars The bond behaviour of the NSM round deformed bars is controlled by splitting tensile stresses in the epoxy cover and the surrounding concrete. Figs. 4a and b illustrate the approach adopted in [11,17] to model this bond behaviour in the plane perpendicular to the bar axis. Two simplifying assumptions are common to these two studies. First, the frictional coecient l (=1/tan c) relating bond shear stresses and internal splitting pressures is constant, although it is known to change during the loading process. Second, the distribution of the radial pressure is uniform, although the pressures on the thicker concrete substrate are higher than those on the thinner cover. De Lorenzis [11] took c to be 45 (i.e. l = 1), whereas Hassan and Rizkalla measured it according to ASTM G115-98 [29]. However, it should be noted that the concept of frictional coecient used in this context is very dierent from that dened by the ASTM specication. The latter pertains to pure frictional properties between materials, depending on material and surface characteristics, whereas the frictional coecient in the bond context is also inuenced by structural variables such as the cover depth and the bar diameter in the case of internal reinforcing bars (e.g. [30]). In the case of NSM reinforcement, even more variables are involved. For steel bars in concrete, Tepfers and Olsson [31] performed ring pull-out tests in order to estimate the angle c of bond stresses at dierent stages of loading. They established curves relating the coecient of friction to the slip for dierent types of bars. Similar measurements for NSM bars would require accurate monitoring of strains in epoxy and concrete transverse to the bar axis. For round deformed bars, Hassan and Rizkalla [19] proposed a model for cover splitting bond failure, based on elastic nite element analysis. They provided two formulae to predict the local bond strengths for the barepoxy and the epoxyconcrete interfaces respectively, and the one corresponding to the smaller local bond strength would control. Depending on how they are computed, these local bond strengths correspond to either the rst cracking of the epoxy cover or the rst cracking of the concrete adjacent to the groove. The two formulae are: smax barepoxy fat l G2 fct l smax epoxyconcrete G1 3 4

and the groove width-to-bar diameter ratios. Graphs for G1 and G2 were developed. The range of the rst ratio examined by them is between 2.00 and 2.50, while that of the second ratio is between 1.50 and 2.50. G1 and G2 range between 0.58 and 1.3 and between 0.5 and 0.72, respectively. Hence, Eq. (4) controls in all cases with values of 0.771.72 times lfct. For a concrete cylinder compressive strength of 2040 MPa, the local bond strength varies between 1.4 and 5.2l, which are very low compared with the test results in Table 2, already with l = 1 instead of 0.5 as originally proposed in Ref. [19]. It should be noted that the so-called local bond strength in this model is the bond shear stress corresponding to the initiation of cracking in the epoxy or in the concrete, whereas the joint can still sustain signicant increments of the applied load between rst cracking and ultimate failure. Also, the distinction between cracking of epoxy and cracking of concrete as two independent modes of bond failure contradicts the experimental evidence that failure in the concrete usually follows cracking of the epoxy. An approximate two-dimensional elastic stress analysis was conducted by De Lorenzis [25] to determine the bond shear stress corresponding to the cracking of the epoxy cover. For computation of an upper and a lower bound to the ultimate radial pressure of the NSM system (and hence to the local bond strength if the frictional coecient is known), dierent possible failure modes were analyzed. A uniform tensile stress distribution along the fracture lines was assumed and justied on the basis of redistribution of cohesive stresses between the crack faces. The experimental values of local bond strength were shown to fall within the computed range. However, no equation for the local bond strength was proposed. 3.4.4. Theoretical models for NSM bars in pre-formed grooves The local bond strength of the epoxy-to-concrete interface for pre-formed grooves was shown to decrease almost linearly with increases in groove size. An equation is given in Ref. [25] that may be used to compute the local bond strength for dierent groove sizes for this case. 3.4.5. Comparison with internal FRP rebars and externally bonded FRP laminates Cosenza et al. [32] presented a summary of bond properties of FRP bars used as internal reinforcement, obtaining an average local bond strength of 2.74 MPa for sand-blasted bars (with a COV of 52%) and of 11.61 MPa for ribbed bars (with a COV of 34%). The local bond strength value for sand-blasted bars is the average of results from Ref. [33], conducted on only one type of bars dierent from those used by De Lorenzis and Nanni [21], hence a direct comparison is not meaningful. The category of ribbed bars referred to by Ref. [32] encompasses a wide variety of material and surface congurations as it covers what are referred to as ribbed bars and spirally wound bars in the present paper, plus some cross-wound

where fat is the tensile strength of epoxy, and G1 and G2 are coecients which were evaluated by nite element analysis and are dependent on the groove depth-to-bar diameter

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

131

bars. The local bond strength values for spirally wound bars from Ref. [32] vary between 4.76 and 18.04 MPa, with an average of 11.9 MPa, which is lower than what has been obtained for NSM systems (Table 2). In bond tests of spirally wound bars as internal reinforcement, signicant interlocking was not evident, but when used as NSM reinforcement, these bars normally produce cover splitting bond failure before the loss of bond at the barepoxy interface. The higher local bond strength for an NSM system can be attributed to a larger resistance at the barepoxy interface than at the barconcrete interface. Nanni et al. [34] conducted bond tests on GFRP ribbed bars of the same type as used by De Lorenzis et al. [26] (although bars in this study had a larger diameter, equal to 12.7 mm), and found a local bond strength of 17 MPa for them as internal reinforcement, which is larger than the value obtained for an NSM conguration. The bond failure of these rebars as internal reinforcement occurred by the shearing-o of the ribs, a mode which was approached but never attained in NSM bond tests as cover splitting bond failure always occurs rst. For FRP laminates externally bonded to concrete, Lu et al. [35] proposed a simple equation for the local bond strength. Assuming an FRP-to-concrete width ratio of 0.5, this equation yields a local bond strength of 1.5 times the tensile strength of concrete, which is signicantly lower than the cover splitting local bond strength of NSM bars. This results from the fact that, in the cover splitting bond

failure mode of an NSM system, the fracture of concrete relates to the larger perimeter (see Fig. 3, mechanisms SP-C1 and SP-C2), while the nominal bond strength is dened using the smaller bar perimeter. In the debonding failure of an externally bonded laminate, the fracture plane has approximately the width of the laminate and the nominal bond strength is based on the same width. 3.5. Local bondslip behaviour Local bondslip curves were deduced from test data by De Lorenzis [25] for dierent types of NSM round bars, and by Sena Cruz and Barros [27] for NSM strips. De Lorenzis [25] reported three dierent types of local bond slip behaviour (types IIII), of which the rst two are shown in Fig. 5. The third type (for sand-blasted round bars) diers from the second in that the abrupt decay from the maximum bond stress to the frictional bond stress level is replaced by a linearly decreasing branch. This third type could be seen as a special case of the second type. The equation proposed by Sena Cruz and Barros [27] has the same form as that of the type I curve shown in Fig. 5. The type I equation seems to reproduce rather accurately experimental curves showing a gradual decrease of local bond stress after the peak. Such a gradual decrease exists when bond failure is at an interface (the epoxyconcrete or the barepoxy interfaces) or by cover splitting generated by ribbed bars with low rib protrusions. In both

Fig. 5. Typical bondslip curves of NSM FRP reinforcement.

132

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

cases, a signicant amount of post-peak friction develops, due to interfacial friction in the rst case and to aggregate interlocking of the cracked concrete in the second case. Conversely, cover splitting failure generated by ribbed bars with high rib protrusions and spirally wound bars has a more brittle nature (for a detailed description see Ref. [26]) with an abrupt decrease in bond stress upon the attainment of the peak value. However, even after the complete loss of the cover, a small amount of residual friction remains because half of the perimeter of the bar is still in contact with the epoxy. The type I equation, where the bond stress tends to zero as the slip approaches innity, cannot reproduce the residual frictional branch. This is only a minor drawback if a 0 is larger than 1, as in such cases the bond stress from the type I equation reaches the frictional plateau only at very large values of slip and the area underneath the bondslip curve is innite. However, if a 0 is smaller than 1, this area becomes nite, and the joint may be predicted to be unable to develop the full tensile capacity of the bar (see next subsection), which contradicts the better behaviour of this type of joint compared with type II joints. More work on local bondslip curves of NSM systems, both experimental and theoretical, is still needed, in particular on the post-peak behaviour which greatly inuences the performance of bars with long embedment lengths such as those used in practice. De Lorenzis et al. [26] commented on the variations of the secant stiness and the shape of the curve with test variables. In nearly all cases, the slip at peak stress was found to be in the range of 0.10.3 mm. Sena Cruz and Barros [27] found a value of 0.25 mm for NSM strips. Based on strain gage and free-end slip readings, Blaschko [16] obtained local bondslip curves of NSM strips, very similar in shape to the type I curves discussed above. However, by plotting the local bondslip curves at dierent measurement locations along the bond length, he noted that the local bond strength tended to be larger close to the free end and smaller close to the loaded end. This was attributed to the inuence of the transverse displacements of the concrete adjacent to the groove, which were also measured in the same tests. In his analytical model, the author adopted a local bondslip curve consisting of an ascending branch dened by a second-degree parabola and a horizontal branch at the local bond strength. This curve was assumed to represent the pure shear stressstrain behaviour of the epoxy. The local bond strength was taken as the shear strength of the epoxy, multiplied by an empirical function of the transverse displacement of concrete. The transverse displacements were predicted with a separate elastic model. Mohrs failure criterion was used to model the eect of local normal stresses, associated with transverse concrete displacements, on the local bond strength. Regression analysis was used to maximize the agreement between measured and predicted strain distributions in the strip along the bond length. In summary, his local bondslip model considers the response of the NSM

bar and the concrete member as a system so that the inuence of edge distance can be eectively reected. However, the model is rather complex to apply as it requires an iterative procedure. Also, it cannot accurately represent the presence of a frictional asymptote in the local bondslip curves obtained from tests and hence it overestimates the bond failure loads of specimens with long bond lengths. For this reason, the author relied on the direct calibration of test results, rather than the output of the bondslip model, to obtain Eq. (1). 3.6. Eective bond length, development length and anchorage length The maximum stress that can be resisted by a bonded joint between an NSM bar and the concrete substrate with a suciently long bond length is given by: r R rmax 2E Gf 5 A with Gf Z
0 1

ssds

where Gf, being the area underneath the bondslip curve, is the fracture energy of the bonded joint, R is the perimeter over which the bond stress acts, A is the cross-sectional area over which the tensile stress acts, and E is the elastic modulus of the material on which the tensile stress is applied [23]. For round bars bonded to epoxy (or to concrete in the case of internal rebars), Eq. (5) reduces to s 8Eb rmax Gf 7 db where Eb is the elastic modulus of the bar, and for laminates bonded to concrete (neglecting the thickness of the adhesive layer), Eq. (5) reduces to s 2Ef Gf rmax 8 tf where Ef and tf are the elastic modulus and thickness respectively of the laminate. If rmax computed by Eq. (5) is below the tensile strength of the reinforcement, its full capacity cannot be developed no matter how long the bond length is. In this case, a value of bond length exists (the eective bond length) at which rmax is developed, and beyond which a further increase in bond length does not produce any benet. The concept of eective bond length has been well established for externally bonded FRP laminates [36,37]. If rmax is larger than the tensile strength of the bar (and in particular for bondslip curves with innite values of Gf), the full capacity of the reinforcement can be developed, and the corresponding value of bond length is usually termed the development length.

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

133

Fig. 6. Idealised bilinear local bondslip model of FRP laminates externally bonded to concrete.

For type I bondslip curves, Gf has an innite or a nite value when a 0 P 1 or a 0 < 1, respectively. For type II responses, if the post-peak frictional branch is unlimited, Gf would have an innite value and the joint would be able to reach the tensile capacity of the reinforcement. In the case of externally bonded laminates, a simplied local bondslip model is shown in Fig. 6. The Gf value for such systems is nite, and its value is usually insucient for the reinforcement to develop its full tensile strength; hence no development length but an eective bond length can be computed. In the case of internal steel or FRP reinforcing bars, Gf is usually innite, due to an unlimited post-peak frictional branch, hence a development length always exists. Available evidence for NSM systems shows that their behaviour is similar to that of internal rebars rather than that of external laminates. The calibrated values of a 0 given in Refs. [25,27] are larger than 1 in all but one case, so the drawback of the type I curve outlined in the previous sub-section has minor practical relevance. Hence, a development length generally exists for an NSM system. However, this statement is based on limited experimental evidence as available bond test results for NSM systems are still limited. The development lengths of NSM round bars in saw-cut grooves computed by De Lorenzis [25] were often impractically long. This was partly due to the need for a better calibration of the post-peak local bondslip response, which displays a signicant scatter and often a quite irregular behaviour particularly in the case of splitting failure. That is, De Lorenzis [25] may have underestimated the ductility of the post-peak response, which led to overestimations of development lengths. For NSM strips, Sena Cruz and Barros [27] predicted a development length of about 90 mm, i.e. less than 10 times the strip height. The failure loadbond length curves (one for each a0e value) proposed by Blaschko [13,16] are composed of a parabolic portion followed by a straight line whose slope is proportional to the post-peak frictional stress. Full development of the tensile capacity of the strip was achieved in the experiments with a relatively short bond length, which is equal to approximately 150 mm (7.5 times the strip height) in absence of edge eects a0e P 150 mm and increases with

a decreasing a0e . The good performance of strips results from the high local bond strength, from the pseudo-ductile post-peak local bondslip response and from the large lateral surface to cross-sectional area ratio. From their tests (see the section on exural strengthening), Hassan and Rizkalla [19] found an eective bond length to exist for NSM round ribbed bars. They computed this eective bond length by assuming a uniform distribution of bond stress equal to the lower of the two local bond strengths given by Eqs. (3) and (4). However, in deriving these equations, bond failure was equated to the imminent cracking of epoxy and concrete, when no signicant bond stress redistribution is expected to have occurred and hence the assumption of a uniform bond distribution is unjustied. Indeed, the assumption of a uniform bond stress distribution contradicts the existence of an eective bond length as the latter implies a limited possibility of bond stress redistribution. These authors did not provide an equation for the maximum tensile stress that can be sustained by the bar, but suggested to obtain it from testing or nite element modelling. The local bondslip curves of NSM FRP bars have been used to determine their bond failure load as a function of the bond length by solving the governing dierential equation [27,21,25]. It is important to note that, while the local bond strength smax dictates the ultimate load of a short NSM FRP-to-concrete joint, the ultimate load of a long joint depends more on the shape of the local bondslip relationship, particularly its post-peak descending branch which controls the ability of the joint to redistribute stresses along its bond length. The shortest bond length required for an FRP bar to resist a given load is herein termed the anchorage length for that load. Note that the anchorage length depends on the load level, and obviously, the anchorage length corresponding to the FRP bar rupture failure load is equal to the development length. In design, it is also important to ensure that the anchorage length provided satises serviceability requirements: the state of the bond stress along the bar should stay within the ascending portion of the bondslip curve and the free end of the bar should not show slip under service loads. Both De Lorenzis [25] and Sena Cruz and Barros [27] provided curves of load at onset of free-end slip versus bond length and outlined a design procedure complying with both ultimate and serviceability limit state requirements. 3.7. Bond test methods The most common types of bond tests used for NSM reinforcement are the direct pull-out test and the beam pull-out test. While detailed descriptions of the various test arrangements can be retrieved from the original papers (see Table 1), some of the issues of concern are discussed below. A number of practical disadvantages exist with beam pull-out tests [14,21]. For example, the specimen size is

134

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

large, especially if long bond lengths are tested; it is dicult to conduct the test in slip-control mode; and it is dicult to visually inspect the behaviour of the joint during loading, especially the initiation and propagation of cracks. Direct pull-out tests overcome the drawbacks of beam pull-out tests mentioned above. The simplest direct pullout test specimen may be composed of a square/rectangular concrete block embedded with an NSM bar on one of the sides, but in this set-up, the NSM bar leads to eccentric loading of the concrete block. The use of two bars on two opposite sides [38] or even four bars on all four sides [3] has been attempted to overcome this problem. The multiplebar specimen has its own problem: any small deviations of the groove/bar positions can induce exural eects, signicantly altering test results. De Lorenzis et al. [23] introduced a C-shaped block where a single NSM bar was placed at the centre of gravity of the block. The set-up performed well, but the specimen dimensions had to be specically designed for each groove depth. This set-up is also not suitable for studying edge eects due to the presence of two thick anges. Based on the above discussions, a direct shear test on a single NSM FRP bar embedded in a concrete block, where the tensile force applied on the bar is reacted by compressive stresses on the concrete block at the loaded end, is probably a good choice that combines simplicity with reliability. A similar test set-up has been popular in studies on externally bonded laminates (e.g. [39]). Blaschko [16] used such a set-up, in which a steel plate was used to provide the reaction to the concrete block. The steel plate had a central hole of 80-mm diameter to avoid reactive stresses on the immediate vicinity of the groove. To minimize the transverse friction generated by the bearing pressure, which could delay the initiation of splitting cracks as generally observed in pull-out tests of steel rebars in concrete, layers of PTFE or similar materials can be placed between the bearing plate and the concrete block. Three main methods are available for obtaining the local bondslip curve of a bonded joint:  Approximate it with the curve relating the average bond stress to the loaded-end slip (or the free-end slip, or the average of the two) from specimens with a short bond length (SBL).  Obtain bond stresses and slips from free-end slip and strain measurements at discrete points along the bond length. This method is usually adopted when long bond length (LBL) specimens are used, as strain gages on a SBL specimen are likely to signicantly aect the bond performance.  Calibrate the unknown parameters in the local bond slip equation, whose form needs to be known or assumed in advance, from loaded-end slip and free-end slip measurements.

Each of the above methods has its advantages and disadvantages. The rst method does not require the use of strain gages, which simplies specimen preparation and does not alter the bond properties between the bar and the epoxy. However, the bond length still needs to be long enough for the specimen to behave as a representative sample (e.g. for ribbed or spirally wound bars, a minimum number of deformations should be included) and to reduce the inuence of end eects. The bondslip curve obtained represents the average performance of the chosen bond length of the specimen. The second method, more onerous and altering to some extent the barepoxy interface, has the advantage that the bond performance over a longer and hence more representative portion of the reinforcement can be studied. A local bondslip curve can be retrieved at each load level, or alternatively at each measurement point, so that more data are obtained from such a test than from one on a SBL specimen. It is worth noting that even for tests conducted with load control on LBL specimens, strain measurements allow the descending branch of the local bondslip relationship to be obtained as a result of stress redistribution over the bond length, although these measurements tend to show a notable scatter, due to damage to the strain gages resulting from slips of the reinforcement. Teng et al. [20] presented tests where strain gauges were sandwiched between two CFRP strips so that the strain gauges were well protected and did not aect the interfacial properties. Wang et al. [40] recently explored the use of ber optic sensors embedded inside FRP bars for the measurement of strains so that the use of strain gauges on the bar surface can be avoided. These and similar techniques for strain measurement are highly desirable for bond tests on LBL specimens to minimize damage to the strain sensors and interference with the interfacial behaviour. The third method has the advantage that it allows the local bondslip curve to be obtained from LBL specimens without the need for strain measurements. The disadvantages are that the form of the bondslip equation must be known a priori, and the accuracy of the deduced equation may be compromised if the assumed form is inappropriate in some way. If many dierent forms of equations are tested to nd the most suitable form, this approach involves a much more onerous process than the other two approaches. In all cases, the specimen must be carefully designed to ensure that the failure mode and load are not signicantly inuenced by the specimen dimensions. An example of inadequate specimen size can be found in Ref. [3], which reports shear fracture failure of the concrete involving the entire specimen cross-section due to its limited section size. When the NSM bar is close to the edge of the concrete block, test results have indicated a strong inuence of the transverse stiness of the concrete on the measured bondslip curve [13], which then also varies

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

135

along the bond length as a result of the variation of the bond stresses. This eect should either be a parameter to be considered or be eliminated by the use of a suciently wide specimen. In addition to section size, the bonded portion of the bar should start at a suitable distance from the loaded end of the concrete block. If this distance is insucient, the behaviour of a specimen with a short bond length may be similar to that of a fastener (see e.g. [38]), or the failure mode of the specimen is unduly inuenced by the cracking of concrete at the loaded end [16]. 4. Flexural strengthening 4.1. Summary of existing work A summary of the existing experimental data is reported in Tables 3 and 4. Some existing studies were conducted on beams strengthened with NSM bars of limited embedment lengths (Table 3). Although such tests were intended to study bond failure mechanisms, they are not pure bond tests as the bond performance is aected by exural cracking. Moreover, the NSM FRP bars in such tests generally extend into the shear spans, where part of the interfacial shear stress is directly dependent on the transverse shear force in the beam. Hassan and Rizkalla [17,19] conducted exural tests on RC beams with NSM CFRP round ribbed bars and strips of varying embedment length. Failure of beams with NSM round ribbed bars occurred by splitting of the concrete cover followed by the complete debonding of the bars in all cases. These authors concluded that the tensile rupture of this type of bars is unlikely to occur, regardless of the embedment length, that the maximum usable strain of these bars should be limited to 0.70.8%, and that the anchorage length should not be shorter than 800 mm. In the case of beams with NSM strips, rupture of the strips occurred when the embedment length was larger than 850 mm. Teng et al. [20] conducted exural tests on RC beams with NSM strips of varying embedment length. As the embedment length increased, the failure mode changed from concrete cover separation starting from the cut-o section, to concrete crushing followed by secondary cover separation close to the maximum moment region. In the beams with the two longest embedment lengths, secondary debonding mechanisms were also observed. All existing test results of strengthened beams, slabs, and columns (Tables 3 and 4) indicate that the NSM reinforcement improved the ultimate load and the load at the yielding of steel reinforcement, as well as the post-cracking stiness. Some test programs included identical beams strengthened with equivalent amounts of FRP provided as either externally bonded or NSM reinforcement. In all cases, the NSM reinforcement performed more eciently, as debonding of the NSM reinforcement occurred at a higher strain or did not occur [17,4143].

One study [41] has compared equivalent amounts of NSM reinforcement provided as round bars or strips. As expected, strips performed better and failed by tensile rupture as opposed to debonding of the round bars, as a result of the higher local bond strength and larger lateral surface to cross-sectional area ratio of NSM strips. 4.2. Failure modes of exurally-strengthened beams The possible failure modes of beams exurally-strengthened with NSM FRP reinforcement are of two types: those of conventional RC beams, including concrete crushing or FRP rupture generally after the yielding of internal steel bars, for which the composite action between the original beam and the NSM FRP is practically maintained up to failure, and premature debonding failure modes which involve the loss of this composite action. Although debonding failures are less likely a problem with NSM FRP compared with externally bonded FRP, they may still signicantly limit the eciency of this technology. The likeliness of a debonding failure depends on several parameters, among which the internal steel reinforcement ratio, the FRP reinforcement ratio, the cross-sectional shape and the surface conguration of the NSM reinforcement, and the tensile strengths of both the epoxy and the concrete. Some researchers [11,10] extended the NSM FRP reinforcement over the beam supports to simulate anchorage in adjacent members. Despite this anchorage, debonding failures can still occur [11]. The beam reported in Ref. [10] failed by FRP rupture, as opposed to debonding observed in an identical beam with the NSM reinforcement terminated away from the supports. Blaschko [16] reported the results of two beam tests: the rst one failed by concrete cover separation starting from the cut-o section but the second beam, which was provided with a steel U-jacket bonded to the cut-o section, failed by the rupture of the FRP strips. In the same study it was observed that fatigue loading to two million load cycles did not aect the residual beam capacity. There is still limited understanding of the mechanics of debonding in beams strengthened with NSM systems. Descriptions of failure modes in the existing literature are often not suciently detailed to understand the progression of the failure process. Based on the available experimental evidence, the possible failure modes of beams exurally-strengthened with NSM FRP reinforcement are classied in Fig. 7 and described below. The interactions between the main failure modes described below and the secondary failure modes are still unclear and deserve further investigations. 4.2.1. Barepoxy interfacial debonding This mode involves interfacial debonding between a bar and the epoxy and has been observed for sand-blasted round bars [44]. This mode correlates well with the failure mode observed in bond tests on the same type of bars (see the previous section). However, unlike in a bond specimen, the

136

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

Fig. 7. Debonding failure modes of NSM bars and strips observed in tests on exurally-strengthened beams: (a) debonding at the barepoxy interface; (b) separation of concrete cover between two cracks in the maximum moment region; (c) separation of concrete cover over a large length of the beam; (d) separation of concrete cover starting from a cuto section; (e) separation of concrete cover along the edge; (f) secondary loss of bond between epoxy and concrete; (g) secondary splitting of the epoxy cover.

epoxy cover in the beam was intersected by exural cracks which facilitated the initiation of longitudinal split-

ting cracks and hence accelerated interfacial debonding (Fig. 7a).

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

137

4.2.2. Concrete cover separation The formation of bond cracks on the sot of the beam has been observed in tests [11,16,20], and these bond cracks are inclined at approximately 45 [20] to the beam axis. Upon reaching the edges of the beam sot, these cracks may propagate upwards on the beam sides maintaining a 45 inclination within the cover thickness, and then propagate horizontally at the level of the steel tension bars. Debonding may next occur in dierent forms, depending on the subsequent evolution of the crack pattern: (a) Bar end cover separation. If the NSM FRP reinforcement is terminated at a signicant distance from the supports, separation of concrete cover typically starts from the cut-o section and propagates inwards [16,20] (Fig. 7d). This mode is similar to the cover separation failure mode observed in RC beams with an externally bonded FRP laminate [37,45,46]. (b) Localized cover separation. Bond cracks within or close to the maximum moment region, together with the pre-existing exural and exural-shear cracks, may isolate triangular or trapezoidal concrete wedges, of which one or more are eventually split o (Fig. 7b). This mode can be identied from photos of failed beams given in Refs. [11,15,20]. (c) Flexural crack-induced cover separation. Separation of the concrete cover occurs almost simultaneously over a long portion of the NSM reinforcement, often involving one of the shear spans and the maximum moment region (Fig. 7c) [15,44]. This mode was observed by De Lorenzis et al. [44] to start from the maximum moment region, whereas the location of initiation was not made clear in Ref. [15]. This mode is similar to the intermediate crack-induced debonding failure mode observed in RC beams with an externally bonded FRP laminate [37,47,48]. (d) Beam edge cover separation. NSM bars located near the edges may generate detachment of the concrete cover along the edges (Fig. 7e). 4.2.3. Epoxyconcrete interfacial debonding For beams with NSM strips of a limited embedment length, Hassan and Rizkalla [17] reported cohesive shear failure in the concrete at the epoxyconcrete interface starting from the cut-o section. Unfortunately no picture of the failed specimens was provided. This mode is believed to be similar to the plate end interfacial debonding failure of RC beams with externally bonded FRP laminate described in Refs. [37,45,46]. 4.2.4. Secondary debonding failure mechanisms Other debonding mechanisms have also been observed. They are herein classied as secondary failure modes and the role they play in the context of debonding failures is still unclear. It has been observed [20] that upon the formation of the bond cracks, the opening-up of these inclined

bond cracks was restrained by the dowel action of NSM reinforcement which in turn tended to cause the detachment of the NSM FRP reinforcement from the sot of the beam. After failure, the prism formed by the CFRP strip and surrounding epoxy was found to retain a thin concrete layer of variable thickness on the sides (Fig. 7f), indicating that a strong epoxyconcrete bond existed. Moreover, localized splitting occurred in the epoxy cover, exposing the internal CFRP strip (Fig. 7g). Similar observations had been reported previously [11]. 4.3. Prediction of ultimate loads and loaddeection behaviour For the safe design of an NSM FRP system for the exural strengthening of an RC beam, the foremost issue is the prediction of the ultimate load. If the failure of a strengthened beam does not involve debonding, then the failure load can be easily predicted using equations developed for externally bonded FRP based on the plane section assumption (e.g. [37]) and with the dierence in position between the two types of reinforcement duly taken into account. Accurate predictions of debonding failure loads are much more challenging. In the past few years, numerous studies have examined debonding failures of beams with externally bonded steel plates and FRP laminates [37,4549]. Research on RC beams exurally-strengthened with NSM FRP has been much more limited. The few theoretical models developed so far are extensions of approaches developed for externally bonded FRP laminates. For instance, Hassan and Rizkalla [17] proposed a theoretical model for NSM strips (valid only for concrete shear failure at the epoxyconcrete interface at the cut-o section) which is an extension of the interfacial stress-based approach proposed by Malek et al. [45]. The model has been compared with a very limited database. Moreover, the failure mode assumed in the model of Hassan and Rizkalla [17] was only observed in the tests by these authors but has not been observed in other tests. Of the reported failure modes, the most critical appears to be the mode of concrete cover separation starting from the maximum moment region. Provided that bars with a reasonable degree of surface deformation are used, failure at the barepoxy interface is unlikely. Moreover, in cases where the NSM reinforcement is needed over the entire length of a member, NSM bars can be easily anchored to adjacent members so that debonding failure at a cuto section can be prevented. For similar reasons, the intermediate crack-induced debonding failure mode has been recognised as the most important mode for RC beams strengthened with externally bonded FRP laminates [47]. For design purposes, the simplest approach to the prediction of debonding is to establish a bond reduction factor for the ultimate tensile strain of the reinforcement. Such an approach is currently adopted by ACI-440 [6] for debond-

138

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

ing of externally bonded laminates. The predictive models in Refs. [47,48] also follow this approach. Alternative approaches [5,50] for this type of debonding failure of externally bonded FRP laminates consider stress gradients in the FRP between two adjacent cracks instead of a simple bond reduction factor [51]. Development of reliable predictive models for debonding failures requires a thorough understanding of the mechanics of debonding failures, and of the qualitative and quantitative roles of relevant variables. The most challenging aspect in tackling this problem appears to be the lack of a direct correlation between the bond failure modes in bond specimens and the debonding failure modes in exurally-strengthened beams. The possible reasons are the presence of exural and exural-shear cracks which alter the bond stress distribution, the curvature of the beam, the dowel action of the FRP bars restraining the opening-up of inclined bond cracks, phenomena which are all absent in a bond specimen. The same problem has been encountered in predicting intermediate crack debonding failures in RC beams with externally bonded FRP laminates, although to a lesser extent [48]. The loaddeection behaviour of beams strengthened with NSM reinforcement can be predicted with reasonable accuracy by the conventional sectional approach neglecting tension stiening and assuming a perfect bond for both steel and NSM reinforcement [11,16]. More rened approaches where tension stiening is taken into account (with dierent laws for un-strengthened and strengthened beams) [15] and where slips of steel and FRP reinforcement are modelled using experimentally determined bondslip equations [52] have delivered more accurate predictions of experimental loaddeection curves. 4.4. Strengthening with pre-stressed NSM FRP NSM FRP bars or strips can be more easily pre-stressed and anchored than externally bonded laminates, so exural strengthening with pre-stressed NSM FRP is a promising technique. Nordin and Taljsten [53] have explored the use of this technique by tensioning NSM bars to 20% of their tensile strength, lling the grooves with epoxy, and releasing the pre-stressing force upon hardening of the epoxy. The expected gains in the cracking load and the stiness of the beam as a result of pre-stressing were achieved, and the failure mode was the tensile rupture of the FRP bars in all cases. The method used by these authors cannot yet be implemented in a real strengthening project as their procedure of tensioning and anchoring the bars requires access to the ends of the beam, which is generally not possible in reality. For this reason, a tensioninganchoring device for NSM bars was proposed by De Lorenzis et al. [54].

5. Shear strengthening 5.1. Summary of existing work The use of NSM FRP reinforcement is also eective in enhancing the shear capacity of RC beams. For this purpose, the bars are embedded in grooves cut on the sides of the member at a desired angle to the beam axis. Only three studies appear to have been published on the use of NSM FRP bars for shear strengthening of RC beams. De Lorenzis and Nanni [55] carried out eight tests on large size T-beams, of which six had no internal stirrups. CFRP ribbed round bars in epoxy-lled grooves were used as NSM shear reinforcement. The test variables included bar spacing, inclination angle and anchorage of the bars in the ange. The NSM reinforcement produced a shear strength increase which is as high as 106% in the absence of steel stirrups, and still signicant in presence of a limited amount of internal shear reinforcement. Barros and Dias [56] tested beams of dierent sizes with no internal stirrups. Some of these beams were strengthened with NSM CFRP strips of dierent inclinations, while the rest were strengthened with equivalent amounts of externally bonded FRP shear reinforcement. The reported strength increases ranged from 22% to 77%, and were in all cases larger than those obtained with externally bonded FRP. Although failure modes were not described, based on the reported loaddeection curves, at least some of the beams are believed to have failed in bending. Nanni et al. [57] reported the test results of a single fullscale PC girder taken from a bridge and shear-strengthened with NSM CFRP strips. The beam failed in exure at a shear force close to the shear resistance predicted by the model given in Ref. [55]. 5.2. Failure modes Two dierent failure modes were identied by De Lorenzis and Nanni [55]. The rst was debonding of the FRP bars by splitting of the epoxy cover and cracking of the surrounding concrete, associated with the diagonal tension failure of concrete (Fig. 8a). This failure mode may be prevented by providing better anchorage of the NSM bars crossing the critical shear crack, by either anchoring the bars in the beam ange or the use of inclined (e.g. 45) bars at a suciently close spacing to achieve a longer total bond length. Once this mechanism was prevented, separation of the concrete cover of the steel longitudinal reinforcement became the controlling failure mode in the tests presented in Ref. [55] (Fig. 8b). Unlike internal steel stirrups, NSM shear reinforcement does not exert a restraining action on the longitudinal reinforcement subjected to dowel forces. These forces, in conjunction with the normal pressures generated by the bond action of the steel longitudinal reinforcement, create considerable tensile stresses in the cover

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

139

Fig. 8. Debonding failure modes of NSM bars observed in tests on shear-strengthened beams: (a) debonding of NSM bars by splitting of epoxy cover; (b) local separation of concrete cover.

which may eventually lead to cover separation failure. This second mode, however, may be attributed to the fact that no or very limited steel stirrups were present in these beams, and is unlikely in beams with a signicant amount of steel stirrups. The most important failure mode is thus debonding of the FRP bars. Although it has not been observed so far, tensile rupture of the NSM reinforcement is another possible failure mode. 5.3. Prediction of ultimate loads The truss analogy was used by De Lorenzis and Nanni [55] to compute the shear capacity of a member strengthened in shear with NSM FRP reinforcement, and in particular the load at diagonal tension failure of concrete involving debonding of the NSM bars. The basic assumption of their approach is that, at the instant of failure, the bond stresses are evenly distributed along the bars crossed by the critical shear crack, and are equal to the local bond strength. This assumption is acceptable if the bondslip behaviour is ductile enough to allow a substantial redistribution of bond stresses along the bars crossed by the shear crack. This basic assumption yields easily the tensile stresses in the bars crossed by the shear crack and hence the corresponding shear force. This approach was shown to compare favourably with test results [55,57]. Further research is obviously needed for the assessment and improvement of the model. Unlike the case of exurally-strengthened beams, existing evidence indicates that the debonding failure mode of NSM bars in shear-strengthened beams is similar to the bond failure mode of the same bars in bond specimens. Further work is needed to conrm this observation as it is important for the modelling of debonding failures of shear-strengthened RC beams. If conrmed, this observation means that local bondslip relationships developed from bond tests can be directly used in predicting debond-

ing failures of RC beams shear-strengthened with NSM FRP bars. With such an approach, the truss model [55] can be easily generalized to the case of debonding failure by incorporating an appropriate bondslip curve instead of the simple ideally plastic curve assumed in the original model. Similarly, local bondslip relationships obtained from bond tests can also be directly used in the numerical modelling of debonding failures. 6. Strengthening of beamcolumn joints Recently, Prota et al. [58] proposed the combined use of FRP laminates and NSM bars for upgrading RC beamcolumn connections. NSM bars were installed on the column prior to wrapping and anchored through the beam (i.e. the bars passed through the beam). This provided additional reinforcement which is fully anchored and eective in the maximum moment region of the column. The presence of FRP wraps prevented the NSM reinforcement from becoming ineective as a result of load reversals. The installation of NSM bars enabled the transition of the failure mode from the column to the shear failure of the joint. In further specimens, additional strengthening was provided to the joint to suppress joint shear failure. For this purpose, the FRP reinforcement was placed in two directions: (a) along the beam axis, NSM FRP bars were provided and transversely conned by single-ply CFRP U-jackets bonded to the beams; (b) along the column axis, a one-ply FRP laminate was provided, and this laminate was terminated below the base of the upper column to simulate the eld condition of the presence of a slab. With this strengthening scheme, failure shifted to the columnjoint interface at the termination of the FRP laminate. The upgrading of the joint zone increased its deformability and hence provided a signicant contribution to the ductility of the system. This study [58] showed that the combination of NSM bars with externally bonded laminates enabled the

140

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

advantages of both techniques to be exploited in a complementary manner. This topic deserves further investigation as similar advantages may be realised by suitable combinations of the two techniques in solving other strengthening problems. 7. Research needs This paper has provided a comprehensive and critical assessment of existing research on the structural behaviour of RC structures strengthened with NSM FRP reinforcement. On the basis of this review, the main research needs in this area for the immediate future are outlined as follows. 7.1. Bond behaviour There is a lack of experimental evidence on the eects of the many variables that are likely to have a signicant eect on the bond behaviour of NSM FRP bars. The number of variables associated with NSM FRP systems is much larger than that for externally bonded FRP systems, but the amount of work available on the former is much more limited than that on the latter. Therefore, extensive further testing is obviously required. Based on existing knowledge and experience, the preferred bond test set-up is believed to be a simple pull-out test where an NSM FRP bar subjected to tension is bonded to a concrete block which is supported at the loaded end. Using such a set-up, eects of factors such as the distance between adjacent grooves (i.e. groove spacing), and the distance between the member edge and the nearest groove (i.e. edge distance) should be examined in the near future. In such bond tests, specimens with either long bond lengths or short bond lengths may be used, and the local bondslip curves obtained from them need to be compared to understand the advantages and disadvantages of using short and long bond lengths. To obtain reliable strain measurements without interference with the interfacial behaviour, the development and application of innovative strain sensors such as ber optic sensors embedded in the FRP bar should also be given due attention. The most important outcome of the bond tests should be local bondslip curves. Further tests are required to assess the existing local bondslip equations, and more importantly to explore the possibility of developing general local bondslip models with parameters expressed as functions of geometry- and material-related properties. To achieve this, the tests should preferably be assisted by numerical modelling of the bond behaviour. For externally bonded FRP laminates, an accurate meso-scale nite element method has recently been developed by Lu et al. [59] which produced numerical results for use with test results in the development of a set of accurate bondslip models [35]. A similar approach should be explored for NSM FRP reinforcement. The model by Lundgren [60] which successfully predicted the bond behaviour of steel

bars in concrete, coupled with appropriate modelling of the epoxyconcrete interface and constitutive modelling of the materials [26], also seems to be a promising approach. Analytical modelling also has a signicant role to play in the modelling of the bond behaviour of NSM FRP bars. In existing bond tests, splitting of the epoxy cover of NSM FRP bars has been identied as an important failure mode. Here, the frictional coecients of dierent deformed bars with dierent groove shapes and dimensions should be measured as a function of the slip in order to further clarify the splitting bond failure mechanism and develop a splitting bond strength model to be used in splitting-critical cases. The approach in Ref. [25] can be considered a rst step towards the denition of a simple splitting bond strength formula accounting for geometric parameters and material properties. For this purpose, the simplifying assumptions of a friction coecient equal to 1 and of a uniform distribution of radial pressures need to be removed, on the basis of experimental measurements of the friction coecient. 7.2. Flexural strengthening Obviously, given the larger number of parameters that can aect the exural behaviour of RC beams with NSM FRP reinforcement, a great deal of further experimental and theoretical work is required. In particular, the debonding failure mechanisms in beams strengthened with NSM reinforcement need to be claried through further testing. The relationship between concrete cover separation and other modes of debonding local to the NSM FRPconcrete joint such as fracture at the epoxyconcrete interface and splitting of the epoxy cover needs further research. Furthermore, the behaviour of pre-damaged beams strengthened with NSM FRP is of signicant practical interest, as cracking and damage to the cover of the steel reinforcement may have a signicant eect on the debonding failure process. The relationship between bond failure mechanisms in bond test specimens and debonding failure mechanisms in exurally-strengthened beams needs to be claried by detailed experimental studies as well as rigorous theoretical modelling. Here, the study of the interaction between exural/exural-shear cracking and bond stresses is of crucial importance. Once this relationship is claried, it will then be possible to develop numerical and analytical models for predicting debonding failures. 7.3. Shear strengthening More tests need to be conducted to further clarify the failure modes of strengthened beams and to evaluate the eects of various signicant factors. More tests are also needed to conrm the applicability of local bondslip models from bond tests in predicting debonding failures of

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143

141

shear-strengthened beams. This conrmation will facilitate the development of accurate numerical and analytical models for RC beams shear-strengthened with NSM FRP. 7.4. Other issues As mentioned earlier in the paper, the combined use of NSM FRP reinforcement in conjunction with externally bonded FRP reinforcement has been found to be eective in strengthening beamcolumn joints [58]. As externally bonded FRP reinforcement alone has met with only limited success in strengthening beamcolumn joints, this combined approach denitely deserves further work. This combined use to take advantages of both techniques should also be explored in solving other strengthening problems. It is widely accepted that pre-stressing the FRP renement before bonding it to concrete structures for strengthening purposes is often desirable, both to improve the serviceability of the structure and to make more ecient use of the FRP material. Pre-stressing externally bonded FRP reinforcement has had little success in practice so far because it is dicult to tension and anchor FRP laminates on site, particularly when they are formed by the wet lay-up process. NSM FRP reinforcement has a much better chance to succeed: NSM FRP bars can be tensioned much more easily, particularly when compared with dry ber sheets, and are much better anchored than externally bonded laminates. The use of cement grout to replace epoxy as the groove ller has been explored by a limited amount of work [9,18,23,26]. There are benets with the use of a cementitious groove ller as discussed earlier in the paper. Research is needed to provide a better understanding of the performance of cement grout as a groove ller and to formulate stronger cementitious groove llers. 8. Concluding remarks Strengthening of structures with NSM FRP reinforcement is a technique that has attracted a considerable attention as a feasible and economic alternative to the technique of strengthening structures with externally bonded FRP reinforcement. The former technique oers some signicant advantages over the latter, including the more ecient use of the FRP material due to a reduced risk of debonding failure and the better protection of the FRP material from external sources of damage. Research on the strengthening of structures using NSM FRP reinforcement started only a few years ago but has by now attracted worldwide attention. A signicant amount of research has been conducted on this emerging technique, particularly on the application of this technique in the strengthening of concrete structures. This paper has provided a detailed and critical review of existing research on the structural behaviour of concrete structures strengthened with NSM FRP reinforcement. This review has shown that

the existing work is still limited in both scope and depth, and many questions remain to be answered before the technique can be widely accepted by practicing engineers. Based on this review, the more urgent research needs have been outlined for NSM FRP strengthening of concrete structures. This paper has been limited to the short-term structural behaviour of concrete structures strengthened with NSM FRP reinforcement, so the use of NSM FRP reinforcement for the strengthening of masonry and timber structures and long-term aspects have not been addressed. Obviously, these additional aspects are equally important, and existing work on them is even more limited. It is thus clear that this emerging strengthening technique poses many challenges as well as opportunities for the international research community. Acknowledgements The work presented in this paper has received nancial support from the Research Grants Council of the Hong Kong SAR (PolyU 5173/04E) and The Hong Kong Polytechnic University (G31 YD 61). The authors are grateful to both organizations for their nancial support. References
[1] Asplund SO. Strengthening bridge slabs with grouted reinforcement. J Am Concr Inst 1949;20(6):397406. [2] Warren GE. Waterfront repair and upgrade, advanced technology demonstration site No. 2: Pier 12, NAVSTA San Diego. Site Specic Report SSR-2419-SHR, Naval Facilities Engineering Service Center, Port Hueneme (CA), 1998. [3] Warren GE. Waterfront repair and upgrade, advanced technology demonstration site No. 3: NAVSTA Bravo 25, Pearl Harbour. Site Specic Report SSR-2567-SHR, Naval Facilities Engineering Service Center, Port Hueneme (CA), 2000. [4] Garrity SW. Near-surface reinforcement of masonry arch highway bridges. In: Proceedings of the 9th Canadian masonry symposium, Fredericton (Canada), CD-ROM, 2001. [5] Fib TG9.3. Externally bonded FRP reinforcement for RC structures. Technical report on the design and use of externally bonded bre reinforced polymer reinforcement (FRP EBR) for reinforced concrete structures. International Federation for Structural Concrete, Lausanne, 2001. [6] American Concrete Institute Technical Committee 440. Guide for the design and construction of externally bonded FRP systems for strengthening concrete structures. ACI 440.2R-02, 2002. [7] De Lorenzis L, Teng JG. Near-surface mounted FRP reinforcement: an emerging technique for structural strengthening. Research Report of Hong Kong Polytechnic University, in preparation. [8] American Concrete Institute Technical Committee 440. Guide for the design and construction of concrete reinforced with FRP bars. ACI 440.1R-03, 2003. [9] Nordin H, Taljsten B. Concrete beams strengthened with CFRP. A study of anchor lengths. In: Proceedings 10th conference on structural faults and repair, London (UK), 2003. [10] Taljsten B, Carolin A, Nordin H. Concrete structures strengthened with near surface mounted reinforcement of CFRP. Adv Struct Eng 2003;6(3):20113. [11] De Lorenzis L. Strengthening of RC structures with near surface mounted FRP rods. PhD Thesis, Department of Innovation Engineering, University of Lecce, Italy, 2002.

142

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143 [33] Alunno Rossetti V, Galeota D, Giammatteo MM. Local bond stress slip relationships of glass bre reinforced plastic bars embedded in concrete. Mater Struct 1995;28(180). [34] Nanni A, Conrad JO, Bakis CE, Uchno J. Material properties of CBar reinforcing bar. Civil Engineering Department, Pennsylvania State University, PA, 1997. [35] Lu XZ, Teng JG, Ye LP, Jiang JJ. Bondslip models for FRP sheets/ plates bonded to concrete. Eng Struct 2005;27(6):92037. [36] Chen JF, Teng JG. Anchorage strength models for FRP and steel plates bonded to concrete. J Struct Eng, ASCE 2001;127(7):78491. [37] Teng JG, Chen JF, Smith ST, Lam L. FRP strengthened RC structures. England: John Wiley & Sons, Ltd; 2002. [38] Yan X, Miller B, Nanni A, Bakis CE. Characterization of CFRP bars used as near-surface mounted reinforcement. In: Proceedings 8th international structural faults and repair conference, Edinburgh (Scotland), CD-ROM version, 1999. [39] Yao J, Teng JG, Chen JF. Experimental study on FRP-to-concrete bonded joints. ComposPart B: Eng 2005;36(2):99113. [40] Wang B, Teng JG, De Lorenzis L, Zhou LM, Ou JP, Jin W, et al. Strain Monitoring of RC Beams Strengthened with Smart NSM FRP Bars, in preparation. [41] El-Hacha R, Rizkalla SH. Near-surface-mounted ber-reinforced polymer reinforcements for exural strengthening of concrete structures. ACI Struct J 2004;101(5):71726. [42] Alkhrdaji T, Nanni A, Chen G, Barker M. Upgrading the transportation infrastructure: solid RC decks strengthened with FRP. Concr Int 1999;21(10):3741. [43] Hassan T, Rizkalla S. Flexural strengthening of prestressed bridge slabs with FRP systems. PCI J 2002;47:7693. [44] De Lorenzis L, Nanni A, La Tegola A. Flexural and shear strengthening of reinforced concrete structures with near surface mounted FRP rods. In: Proceedings ACMBS III, Ottawa (Canada), 2000. p. 5218. [45] Smith ST, Teng JG. FRP-strengthened RC beams. I: review of debonding strength models. Eng Struct 2002;24(4):38595. [46] Smith ST, Teng JG. FRP-strengthened RC beams. II: assessment of debonding strength models. Eng Struct 2002;24(4):397417. [47] Teng JG, Smith ST, Yao J, Chen JF. Intermediate crack-induced debonding in beams and slabs. Constr Build Mater 2003;17(67): 44762. [48] Lu XZ, Teng JG, Ye LP, Jiang JJ. Intermediate crack debonding in FRP strengthened RC beams: FE analysis and strength model, submitted for publication. [49] Oehlers DJ, Seracino R. Design of FRP and steel plated RC structures: retrotting beams and slabs for strength, stiness and ductility. Elsevier; 2004. [50] JSCE. Recommendations for upgrading of concrete structures with use of continuous ber sheets. Japan Society of Civil Engineers, 2001. [51] Yao J, Teng JG, Lam L. Experimental study on intermediate crack debonding in FRP-strengthened RC exural members. Adv Struct Eng 2005;8(4):36596. [52] Aiello MA, De Lorenzis L, Micelli F. Loaddeection behavior of RC beams strengthened with near-surface mounted FRP bars. In: Proceedings CCC2003, Cosenza (Italy), September 2003. [53] Nordin H, Taljsten B. Concrete beams strengthened with prestressed near surface mounted reinforcement. In: Proceedings FRPRCS-6, Singapore, 2003. p. 107786. [54] De Lorenzis L, Micelli F, La Tegola A. Passive and active nearsurface mounted FRP rods for exural strengthening of RC beams. In: Proceedings ICCI-02, San Francisco (CA), CD-ROM version, 2002. [55] De Lorenzis L, Nanni A. Shear strengthening of reinforced concrete beams with NSM ber-reinforced polymer rods. ACI Struct J 2001;98(1):608. [56] Barros JAO, Dias S. Shear strengthening of reinforced concrete beams with laminate strips of CFRP. In: Proceedings CCC2003, Cosenza (Italy), 2003. p. 28994.

[12] Parretti R, Nanni A. Strengthening of RC members using nearsurface mounted FRP composites: design overview. Adv Struct Eng 2004;7(6):46983. [13] Blaschko M. Bond behaviour of CFRP strips glued into slits. In: Proceedings FRPRCS-6. Singapore: World Scientic; 2003. p. 20514. [14] Sena Cruz JM, Barros JAO. Bond behavior of carbon laminate strips into concrete by pull-out bending tests. In: Proceedings of the international symposium Bond in concrete from research to standards, Budapest (Hungary), 2002. p. 61421. [15] Barros JAO, Fortes AS. Flexural strengthening of concrete beams with CFRP laminates bonded into slits. Cement Concr Compos 2005;27(4):47180. [16] Blaschko M. Zum tragverhalten von betonbauteilen mit in schlitze eingeklebten CFK-lamellen. Bericht 8/2001 aus dem Konstruktiven Ingenieurbau, TU Munchen, 2001. 147 pp [in German]. [17] Hassan T, Rizkalla S. Investigation of bond in concrete structures strengthened with near surface mounted carbon ber reinforced polymer strips. ASCE J Compos Constr 2003;7(3):24857. [18] Carolin A, Taljsten B. Behaviour of concrete beams strengthened with near-surface mounted reinforcement, NSMR. In: Proceedings ACIC2002, Southampton (UK), 2002. p. 17784. [19] Hassan T, Rizkalla S. Bond mechanism of near-surface-mounted ber-reinforced polymer bars for exural strengthening of concrete structures. ACI Struct J 2004;101(6):8309. [20] Teng JG, De Lorenzis L, Wang B, Rong L, Wong TN, Lam L. Debonding failures of RC beams strengthened with near-surface mounted CFRP strips. J Compos Constr, ASCE 2006;10(2):92105. [21] De Lorenzis L, Nanni A. Bond between NSM ber-reinforced polymer rods and concrete in structural strengthening. ACI Struct J 2002;99(2):12332. [22] Al-Zahrani MM. Bond behavior of ber reinforced plastic (FRP) reinforcements with concrete. PhD Thesis, Department of Civil and Environmental Engineering, The Pennsylvania State University, University Park (PE), 1995. p. 263. [23] De Lorenzis L, Rizzo A, La Tegola A. A modied pull-out test for bond of near-surface mounted FRP rods in concrete. ComposPart B: Eng 2002;33(8):589603. [24] Tepfers R. A theory of bond applied to overlapped tensile reinforcement splices for deformed bars. Publication 73:2, Division of Concrete Structures, Chalmers University of Technology, Gothenburg (Sweden), 1973. 328 pp. [25] De Lorenzis L. Anchorage length of near-surface mounted FRP bars for concrete strengthening analytical modeling. ACI Struct J 2004;101(3):37586. [26] De Lorenzis L, Lundgren K, Rizzo A. Anchorage length of nearsurface mounted FRP bars for concrete strengthening experimental investigation and numerical modeling. ACI Struct J 2004;101(2):26978. [27] Sena Cruz JM, Barros JAO. Modeling of bond between near-surface mounted CFRP laminate strips and concrete. Comput Struct 2004;82(1719):151321. [28] MacGregor JC. Reinforced concrete mechanics and design. 3rd ed. Upper Saddle River (NJ): Prentice-Hall, Inc.; 1997, 939 pp. [29] ASTM G115-98. Standard guide for measuring and reporting friction coecients. ASTM International, West Conshohocken (PA), 2001. 11 pp. [30] van der Veen C. Cryogenic bond stressslip relationship. Delft University of Technology, Doctoral Thesis, Delft (The Netherlands), 1990. 111 pp. [31] Tepfers R, Olsson PA. Ring test for evaluation of bond properties of reinforcing bars. In: Proceedings of the international conference on bond in concrete from research to practice, Riga, Latvia, vol. 1, 1992. p. 8999. [32] Cosenza E, Manfredi G, Realfonzo R. Behavior and modeling of bond of FRP rebars to concrete. ASCE J Compos Constr 1997;1(2):4051.

L. De Lorenzis, J.G. Teng / Composites: Part B 38 (2007) 119143 [57] Nanni A, Di Ludovico M, Parretti R. Shear strengthening of a PC bridge with NSM CFRP rectangular bars. Adv Struct Eng 2004;7(4):97109. [58] Prota A, Nanni A, Manfredi G, Cosenza E. Selective upgrade of underdesigned RC beamcolumn joints using CFRP. ACI Struct J 2004;101(5):699707. [59] Lu XZ, Ye LP, Teng JG, Jiang JJ. Meso-scale nite element model for FRP sheets/plates bonded to concrete. Eng Struct 2005;27(4):56475. [60] Lundgren K. Bond between ribbed bars and concrete: Part 1. Modied model. Mag Concr Res, in press.

143

[61] Crasto A, Kim R, Ragland W. Private communication, University of Dayton Research Institute, 1999. [62] Yost JR, Gross SP, Dinehart DW, Mildenberg J. Near surface mounted CFRP reinforcement for the structural retrot of concrete exural members. In: Proceedings ACMBS-IV, Calgary (Canada), July 2004, CD-ROM. [63] Arduini M, Romagnolo M, Camomilla G, Nanni A. Inuence of concrete tensile softening on the performance of FRP strengthened RC beams: experiments. In: Proceedings ACMBS-IV, Calgary (Canada), July 2004, CD-ROM.

Das könnte Ihnen auch gefallen