You are on page 1of 18

Construction and Building Materials 132 (2017) 94–111

Contents lists available at ScienceDirect

Construction and Building Materials


journal homepage: www.elsevier.com/locate/conbuildmat

Review

Fabric-reinforced cementitious matrix: A promising strengthening


technique for concrete structures
Oluwafunmilayo Awani, Tamer El-Maaddawy ⇑, Najif Ismail
United Arab Emirates University, United Arab Emirates

h i g h l i g h t s

 Previous studies on strengthening of RC structures with FRCM are critically reviewed.


 Factors affecting the performance of FRCM-strengthened RC elements are highlighted.
 Research gaps are identified.
 Directions for future research are outlined.

a r t i c l e i n f o a b s t r a c t

Article history: Fabric-reinforced cementitious matrix (FRCM) systems have recently been introduced in the construction
Received 26 June 2016 industry as a viable alternative strengthening material, to circumvent problems associated with fiber-
Received in revised form 11 November 2016 reinforced polymers (FRP). They are made of fabric grids and a cementitious agent which serves as matrix
Accepted 25 November 2016
and binder. The cementitious matrix used in FRCM systems has higher thermal capacity and better com-
patibility with the concrete substrate compared to those of the epoxy resin used in FRP. The use of FRCM
composites for strengthening and repair of reinforced concrete structures, though relatively recent, is
Keywords:
gradually gaining popularity as an alternative to FRP. This paper presents a critical review of existing
Bond
Cementitious matrix
research on structural strengthening with FRCM composites, identifies gaps in knowledge, and outlines
FRCM directions for future research.
Fabric Ó 2016 Elsevier Ltd. All rights reserved.
Fiber
Flexural
Mortars
Shear
Textile

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
2. Bond behaviour of FRCM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
2.1. Test methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
2.2. Failure modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
2.3. Factors affecting bond behaviour of FRCM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
2.3.1. Number of FRCM layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
2.3.2. Fabric type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
2.3.3. Anchorage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
2.3.4. Bond width . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
2.3.5. Bond length. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
2.3.6. Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
2.4. Bond-slip (s-s) models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

⇑ Corresponding author at: United Arab Emirates University, Al Ain, P.O. Box 15551, United Arab Emirates.
E-mail addresses: more_funmi@yahoo.com (O. Awani), tamer.maaddawy@uaeu.ac.ae (T. El-Maaddawy), najif@uaeu.ac.ae (N. Ismail).

http://dx.doi.org/10.1016/j.conbuildmat.2016.11.125
0950-0618/Ó 2016 Elsevier Ltd. All rights reserved.
O. Awani et al. / Construction and Building Materials 132 (2017) 94–111 95

3. Flexural strengthening with FRCM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100


3.1. Failure modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3.1.1. Concrete cover separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3.1.2. Debonding at fabric-matrix interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3.1.3. Debonding at concrete-matrix interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3.1.4. Slippage of fabric roving within matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3.1.5. Rupture of fabric within matrix. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3.2. Factors affecting the flexural capacity of FRCM-strengthened members . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3.2.1. Number of FRCM layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3.2.2. Fabric type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
3.2.3. Mortar type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
3.2.4. FRCM configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
3.2.5. Anchorage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.2.6. Concrete strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.2.7. Corrosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.2.8. Tensile steel reinforcement ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
4. Shear strengthening with FRCM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
4.1. Failure modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
4.1.1. Cover separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
4.1.2. Debonding at concrete-matrix interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
4.1.3. Slippage of fabric roving within the matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.1.4. Rupture of fabric within the matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.2. Factors affecting the shear capacity of FRCM-strengthened members. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.2.1. Number of FRCM layers (reinforcement ratio) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.2.2. Fabric type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.2.3. Mortar type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.2.4. FRCM configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.2.5. Mechanical anchorage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
4.2.6. Concrete strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
4.2.7. Stirrup spacing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
4.2.8. Mortar thickness. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5. Column confinement with FRCM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5.1. Failure modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5.1.1. Fabric rupture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5.1.2. Debonding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.1.3. Fiber slippage within rovings. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.1.4. Concrete crushing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.2. Factors affecting the axial capacity of FRCM-strengthened members . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.2.1. Number of FRCM layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.2.2. Fabric type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.2.3. Mortar type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.2.4. FRCM configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.2.5. Concrete strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.2.6. Tie spacing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.2.7. Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.2.8. Section size/shape. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.2.9. Layout of longitudinal steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.2.10. Corner radius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.2.11. Load eccentricity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.2.12. Loading protocol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
6. Conclusions and recommendations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
6.1. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
6.2. Recommendations for future studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
Notations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

1. Introduction bond surface/line. This often leads to delamination of the compos-


ite from the substrate [1–8]. FRP composites also have poor perfor-
Since the early 1980’s, fiber-reinforced polymer (FRP) compos- mance at elevated temperatures and thermal incompatibility with
ites have gradually gained popularity as viable strengthening the substrate. They are difficult to apply on humid surfaces and at
materials for reinforced concrete (RC) structures. FRP composites low temperatures. Fabric-reinforced cementitious matrix (FRCM)
possess certain characteristics, which make them a viable alterna- systems, also referred to as textile-reinforced mortar (TRM),
tive to the traditionally used strengthening systems. These charac- mineral-based composite (MBC) or textile-reinforced concrete
teristics include corrosion resistance, high strength-to-weight (TRC) systems, have recently been introduced in the construction
ratio, and high versatility. FRPs consist of fibers embedded in a industry as an alternative to circumvent the problems associated
polymer matrix, usually epoxy resin. The epoxy resin used in FRPs with FRPs [9]. FRCM composites consist of fabric grids (made of
is incompatible with the concrete substrate, resulting in a critical fibers such as carbon, glass, etc.) and a cementitious agent (mortar)
96 O. Awani et al. / Construction and Building Materials 132 (2017) 94–111

that serves as matrix and binder. The matrix used in FRCM com- front’ or strain peak could develop close to the loaded end then
posites has higher thermal capacity compared to that of the epoxy shifts towards the unloaded end with further increase in load [10].
resin used in FRPs, and is more compatible with the concrete sub-
strate. The inclusion of mortar in the system rather than epoxy 2.1. Test methods
improves the heat/fire resistance and the compatibility with the
concrete substrate [9]. Two test methods are typically employed to investigate the
FRCM composites are applied in a similar manner as FRP com- bond behaviour of FRCM: single-lap shear tests, and double-lap
posites for the strengthening and repair of RC structures. They shear tests. In single-lap shear tests (Fig. 3), the FRCM is applied
are usually applied by the hand lay-up method, which involves to one surface of the substrate (concrete prism) with a length of
the application of a layer of mortar, followed by a layer of fabric bare fabric extending beyond the bonded area. The full length of
grid, and a second mortar layer. This process is repeated till the the bare fabric or just its end (away from the bonded area) is usu-
required number of layers is reached. The bond between the fabric ally sandwiched between steel plates using epoxy. A pull force is
and the mortar is achieved by mechanical interlock, formed as a then applied to the fabric while the concrete substrate remains
result of the mortar penetrating through the grid spaces. It is fixed [11,13,16,18].
important that each mortar layer is applied while the previous In double-shear tests, FRCM strips are bonded on opposite sides
layer is still wet. In flexural strengthening, FRCMs are typically of the concrete substrate and a force is applied simultaneously to
applied externally at the bottom surface of RC beams and slabs. both strips. The force applied to the FRCM can be achieved by
In shear strengthening, they are applied at the lateral sides of RC bonding the FRCM to two concrete prisms, and pulling the prisms
beams, in a side-bonded, U-shape, or full-wrapping configuration. apart using rods inserted through their middle (Fig. 4), or by apply-
The use of FRCM composites for strengthening and repair of RC ing a push force between both prisms as shown in Fig. 5.
structures, though relatively recent, is gradually gaining popularity Bond failure of the FRCM systems typically occurs at the fabric-
as an alternative to FRPs. Existing literature on the use of FRCM matrix interface. In the single-lap test, the failure load can be mea-
composites is much more limited than FRP composites. This paper sured directly with more accuracy than in double-lap tests. In
presents a critical review of existing research on structural double-lap tests, unsymmetrical debonding could occur due to
strengthening with FRCM composites, identifies gaps in knowl- the prisms rotating relative to their initial positions, leading to
edge, and outlines directions for future research. A summary of one FRCM strip debonding before the other. Load re-distribution
available literature on bond behaviour of FRCM, flexural strength- occurs in double-lap shear tests after debonding occurs, and may
ening, shear strengthening, and column confinement with FRCM even happen before failure due to eccentricity or inconsistencies
are discussed in the following sections. in bonding characteristics of the two composite strips [14,17]. As
a result of this, direct calculation of the applied load to each strip
by halving the total load may be inaccurate, especially in describ-
2. Bond behaviour of FRCM ing the post-peak response. Sneed et al. [17] proposed a methodol-
ogy to determine the load applied to each FRCM strip using LVDT
In order to fully maximize the potential of FRCM strengthening measurements, the angle of rotation of the prisms, and the hori-
materials, a sufficient understanding of their bond behaviour at the zontal translation of the prisms.
fabric-matrix interface and with the concrete substrate is essential. Single-lap shear test setup is recommended for FRCM-concrete
FRP materials exhibit a linear stress-strain response, however this bond tests due to its simplicity, and as double-lap shear setup
is not the case with FRCM composites. FRCM composites have a bi- tends to estimate the lower bound of the composite’s bond capac-
linear stress-strain behaviour, with a change in slope representing ity [17].
cracking [9,10]. Fig. 1 compares the typical stress-strain response
of FRP with that of FRCM. Table 1 gives a summary of previous 2.2. Failure modes
studies [11–18] on the bond behaviour of FRCM. Results of these
studies are discussed in the following sections. Failure modes observed in FRCM bond tests are presented in
The strains in FRCM composites tend to be higher at the loaded Table 1. Debonding at the fabric-matrix interface is the most
end and decrease towards the unloaded end as the load increases prevalent bond failure mode of FRCM systems. It is characterized
[10,12,14,16]. This behaviour is also typical in FRP strengthening. by transverse cracks which initiate near the loaded end of the spec-
Strain distribution along the bond length of FRCM in bond speci- imen and propagates towards the unloaded end as the applied load
mens tested by the first author is shown in Fig. 2 [19]. A ‘strain increases. The transverse cracks extend from a horizontal crack
which forms at the fabric-matrix interface, as shown in Fig. 6.
The fabric-matrix debonding mode of failure is accompanied by
slippage and deformation of the longitudinal fiber bundles [12–
14]. This was clearly observed in specimens tested by Sneed
et al. [13], where the longitudinal fiber bundles pulled out of the
matrix at the loaded end.
In some cases, bond failure of FRCM is characterized purely by
slippage of the fabric inside the matrix as exhibited by specimens
with one FRCM layer tested by Ombres [16]. Depending on the type
of matrix used, inconsistencies could be noted in the bond between
the matrix and the concrete, resulting in debonding of the FRCM
system from the concrete substrate [16]. This highlights the impor-
tance of the substrate surface preparation prior to application of
the FRCM to ensure adequate bond at the concrete-matrix
interface.
Another mode of failure observed in the literature is rupture of
the fabric in the matrix. This was observed by Awani et al. [14]
Fig. 1. Typical tensile stress-strain response of FRCM and FRP coupons. in FRCM-strengthened specimens using a matrix that exhibited
Table 1
Summary of previous studies on bond behaviour of FRCM.

Reference
Hashemi and Al- D’Ambrisi et al. Sneed et al. Awani et al. [14] D’Antino et al. Ombres [16] Sneed et al. [17] D’Antino
Mahaidi [11] [12] [13] [15] et al. [18]
No. of specimens 5 12 58 18 15 24 16 (+ 58 from Sneed 20
et al. [13])
Test parameters Bond length, Bond length, Bond length, Bond length, bond Loading Bond length, bond width, number Bond length, bond Length of
anchorage, fabric number of layers bond width width, matrix type frequency, load of layers, temperature width, test method prism
type range
Test method Single-lap shear Double-lap shear Single-lap Double-lap shear Single-lap shear Single-lap shear Double-lap shear Single-lap
shear shear

O. Awani et al. / Construction and Building Materials 132 (2017) 94–111


Loading protocol Monotonic Monotonic Monotonic Monotonic Cyclic-fatigue Monotonic Monotonic Monotonic
Concrete Specimen Prism Prism Prism Prism Prism Prism Prism Prism
shape
Width (mm) 75 100 125 150 125 N/A 125 125
Length (mm) 245 N/A 375 150 375 N/A 375 375, 510
Comp. 38 (cyl.) N/A 42.5 (cyl.) 55 (cube) 42.5 (cyl.) N/A 33.5 (cyl.) 33.5, 42.5
strength (cyl.)
(MPa)
Details of Fabric Fabric type Carbon PBO PBO Carbon PBO PBO PBO PBO
FRCM ff (MPa) N/A 5213/53911 3015 3800 3010 5213/53911 3014 3015
Ef (GPa) N/A 271/2732 206 230 206 271/2732 206 206
tf (mm) 0.176 0.046/0.0113 0.092 N/A 0.092 0.046/0.0223 0.092 0.092
bf (mm) 50 N/A 34, 43, 60 75, 100, 150 60 50, 70 34, 60, 80, 100 60
L (mm) 100, 180 N/A 100, 150, 200, 75, 100, 150 330 150, 200, 250 100, 200, 250, 330 330
250, 330
efu (%) N/A 1.92/1.974 1.45 1.42 1.45 1.92/1.97 1.45 1.45
Mortar fc,m (MPa) 65 (cube) 16.1+ 27.9 (cyl.) 65, 69 (cube) 28.4 (cyl.) 30.4 (cube) 28.4 (cyl.) 28.4 (cyl.)
ft,m (MPa) 5.7* 2.6* 3.6** 3.4, 3.5** 3.5** N/A 3.5** 3.5**
Em (GPa) N/A 6.1 N/A 25.4, 40.7 N/A 6.1 N/A N/A
tm (mm) N/A N/A 4 4 4 4 4 4
Results Failure mode DB2 DB2, FS DB2, FS FR, FS, DB2 FR FS, DB2, DB3 FS, DB2 FS, DB2
Failure load N/A 5.2–15.6 0.97–7.12 8.4–38.7 2.3–3.3 1.6–8.9 3.3–10.6 5.2–8.3
(kN)
smax (mm) N/A 0.26–1.17 N/A N/A N/A 0.22–0.63 N/A N/A
emax (%) N/A 7.31–12.96 N/A N/A N/A 0.41–0.99 N/A N/A
Le (mm) N/A 250–300 250–330 N/A N/A 150–200 260 N/A

cyl. = cylinder compressive strength of concrete; cube = cube compressive strength; PBO = poly-p-phenylenebenzobisoxazole.
Failure modes: DB2 = debonding at fabric-matrix interface; DB3 = debonding at concrete-matrix interface; FR = fabric rupture; FS = fabric slippage; bf = width of fabric; Ef = Young’s modulus of fiber; Em = Young’s modulus of
mortar; fc,m = compressive strength of mortar; ff = tensile strength of fiber; ft,m = tensile strength of mortar; L = bond length; Le = effective bond length; tf = nominal thickness of fabric; tm = thickness of one layer of mortar;
smax = relative displacement between fabric and matrix at failure; efu = ultimate tensile strain of fiber; emax = fiber strain at failure.
*
Determined from flexure test.
**
Determined from split cylinder test.
+
Determined using halved prisms from flexure test.
1
Tensile strength of fibers in longitudinal/transversal directions.
2
Young’s modulus of fibers in longitudinal/transversal directions.
3
Nominal fiber bundle thickness in longitudinal/transversal directions.
4
Ultimate tensile strain of fiber in longitudinal/transversal directions.

97
98 O. Awani et al. / Construction and Building Materials 132 (2017) 94–111

Fig. 2. Surface strain distribution along bond length of FRCM in bond specimens
tested by the first author [19].

Fig. 3. Typical single-lap shear test setup.

Fig. 5. Double-lap push shear test adopted previously by the first author.

failure mode from fabric slippage to delamination at the fabric-


matrix interface.

2.3.2. Fabric type


Hashemi and Al-Mahaidi [11] reported that fabric textiles (with
spacing between tows) had better bond with the cementitious
mortar than fabric sheets. This was ascribed to penetration of the
mortar through the fabric, which resulted in a mechanical interlock
thus improving the bond between the fabric and the mortar. The
bond capacity of a specimen with a fabric textile was 30% more
Fig. 4. Typical double-lap pull shear test setup.
than that of a corresponding specimen with a fabric sheet.

irregular bonding behaviour with the fabric. Rupture of fabric in 2.3.3. Anchorage
the matrix was also reported in specimens subjected to cyclic load- Hashemi and Al-Mahaidi [11] investigated the effect of anchor-
ing. D’Antino et al. [15] speculated that the fabric was damaged age consisting of a 30  20 mm groove created in the concrete sub-
due to friction between the fabric and matrix in the course of the strate in which the fabric was held using two FRP strands. The
repeated load cycles. concept of the anchorage system used is illustrated in Fig. 7. The
presence of anchorage improved the bond between the FRCM
and the concrete substrate, delayed debonding of the fabric, and
2.3. Factors affecting bond behaviour of FRCM thus resulted in higher bond capacity of the specimens. The bond
capacity of a specimen with anchorage was 25% higher than that
Different parameters affecting the bond behaviour of FRCM of a similar specimen without anchorage. In another pair of speci-
were investigated in previous studies. The effects of these param- mens, an increase in bond capacity corresponding to about 15%
eters are discussed in the following sections. was observed due to anchorage. The increase in the bond capacity
of the latter pair of specimens was less significant than that of the
former because the debonding cracks, in the latter pair, did not
2.3.1. Number of FRCM layers extend to the anchorage zone.
D’Ambrisi et al. [12] noted that specimens with two FRCM lay-
ers experienced lower strains in the fabric than those of specimens 2.3.4. Bond width
with one FRCM layer. The debonding fabric strain in specimens Sneed et al. [13,17] reported that varying the bond width of the
with two FRCM layers was approximately 85% of that of specimens FRCM composites tended to have no effect on their bond perfor-
with one FRCM layer. Ombres [16] also observed a reduction in mance. This was assumed to be as a result of independent action
fabric strain with an increase in number of layers, and a shift of of the longitudinal fiber bundles [13]. Ombres [16], however,
O. Awani et al. / Construction and Building Materials 132 (2017) 94–111 99

Fig. 6. Typical bond failure at fabric-matrix interface.

slip between the fabric and matrix, since there was no slippage
between the mortar and the concrete substrate.
The perimeter of the bond surface is important in the calibra-
tion of the bond-slip relationship, and is defined as pf = 2bf for
one fabric layer, where bf is the width of the fabric layer. With n
fabric layers, however, two assumptions of pf can be made: (i) all
fabric layers are considered as one homogenous layer, having a
thickness of n  tf, and pf = 2bf; or (ii) all fabric layers are consid-
ered as independent, therefore pf = 2nbf, where n = number of fab-
Fig. 7. Concept of FRCM anchorage system with groove in concrete. ric layers. Experimental results of D’Ambrisi et al. [12,20] indicated
that the effective perimeter of the bond surface for multiple FRCM
layers ranged between assumption (i) and assumption (ii). For this
reported an effect for the bonded width on the bond-slip response. reason, a bond effectiveness factor [k = k(n)] was introduced by
With similar bond length, specimens with a bond width of 50 mm D’Ambrisi et al. [20], thus pf is expressed as:
showed higher bond stress values at failure compared with speci-
mens with a bond width of 70 mm. This was attributed to the non- pf ¼ 2nbf  kðnÞ kð1Þ ¼ 1; and 0 6 kðnÞ 6 1 ð1Þ
uniform distribution of fibers across the bonded width. The assumptions (i) and (ii) are determined by Eq. (1), assuming
k(n) = 1/n and k(n) = 1, respectively. The bond-slip laws given in
2.3.5. Bond length Eqs. (2) and (3) provided the most accurate predictions of the
The bond capacity of FRCM specimens increases with an experimental results of D’Ambrisi et al. [12,20], where s = slip (rel-
increase in the bond length. By plotting the bond capacity against ative displacement between fabric and matrix); sf = final slip when
the bonded length, with sufficient bonded lengths, an effective value of bond stress is zero; A, sr, a, b, c = curve-fitting parameters;
length can be clearly observed beyond which the bond capacity s = bond stress; and so = initial finite value of bond shear stress.
remains constant. The effective bond length (Le) was found to be  
in the range of 250–300 mm, 250–330 mm, and 150–200 mm, as s
sðsÞ ¼ ½s0 þ Aðeas  ebs Þ  1  0 6 s 6 sf ð2Þ
reported by D’Ambrisi et al. [12], Sneed et al. [13], and Ombres sf
[16], respectively. Awani et al. [14] reported a decrease in maxi-
"  #
mum bond stress at the concrete-matrix interface with an increase s
2
2s
in the bonded length, for the same bonded area. sðsÞ ¼ s0 ½Aðs þ sr Þa þ 1  ecs  þ1 0 6 s 6 sf ð3Þ
sf sf

2.3.6. Temperature The unknown parameters of these equations were evaluated


Bond specimens in Ombres [16] were exposed to temperatures such that the adopted values minimized the difference between
of 50 °C and 100 °C for a period of 8 h, and allowed to cool before the experimental results and analytical predictions. The values of
testing at an ambient temperature of 20 °C. A decrease in failure these parameters are given in Table 2. A typical s-s curve of FRCM
load with an increase in temperature was reported. With two specimens is shown in Fig. 8. Ombres [16] calibrated a bond-slip
FRCM layers, the failure load decreased by 28% and 38% with relation assuming pf = 2nbf using a procedure similar to that used
heat-conditioning at 50 °C and 100 °C, respectively. Temperature by D’Ambrisi et al. [20]. The curve proposed by Ombres [16] was
conditioning at 100 °C for specimens with one FRCM layer resulted more conservative. The proposed models are influenced by the
in a 36% reduction in the failure load, whereas no strength reduc- experimental results used in the calibration, and may not be accu-
tion was observed at 50 °C. The reduction in the failure load of the rate for other FRCM strengthening systems.
heat-conditioned specimens could be due to a change in the failure Jung et al. [21] modified a model developed earlier by Teng et al.
mode. Specimens which were not heat-conditioned failed by slip- [22] to predict the bond stress in FRCM at debonding, rdeb, (Eq. (4)).
page or debonding at the fabric-matrix interface, while the heat- The modified model defines the total nominal thickness of the fab-
pffiffiffi
conditioned specimens failed mostly by debonding at the fabric- ric as t f ;total ¼ t f  n , as suggested by D’Ambrisi et al. [12,20].
matrix or concrete-matrix interfaces. vffiffiffiffiffiffiffiffiffiffiffiffiffi
u qffiffiffiffi0
uE f
t f
rdeb
c
2.4. Bond-slip (s-s) models ¼ ab bf bL pffiffiffi ð4Þ
tf n
Bond failure of FRCM typically occurs at the fabric-matrix inter- qffiffiffiffiffiffiffiffiffiffiffiffiffi
2bf =bw pL (if L < L ).
face with little or no slippage of the matrix relative to the concrete where bf ¼ 1þb =bw
, and bL ¼ 1 (if L P Le) or sin 2L e
e
f

substrate, indicating adequate bond with the substrate [14,16]. The parameter ab was taken as 0.729 based on regression anal-
D’Ambrisi et al. [12,20] used their experimental results to develop ysis. This model was used to predict the ultimate loads at the onset
bond-slip laws for FRCM. The displacement was assumed to be the of debonding at concrete-matrix interface of the beam specimens
100 O. Awani et al. / Construction and Building Materials 132 (2017) 94–111

Table 2 dinal crack at the level of the fabric extending from a flexural crack
Parameters of bond-slip relationships. in the maximum moment region. The longitudinal crack gradually
Parameters D’Ambrisi et al. [20] Ombres [15] propagates towards the ends of the FRCM strengthening layer
Eq. (2) Eq. (3) Eq. (2) (Fig. 9b). This failure mode was observed in Hashemi and Al-
pf = 2nbf  k(n) pf = 2nbf  k(n) pf = 2bf pf = 2nbf Mahaidi [11], El-Maaddawy and El Refai [26], D’Ambrisi and
A 0.92 0.25 1.15 0.7
Focacci [27], and Ebead et al. [28].
a 8.94 4.55 0.007 0.005
sf 1.81 229 0.47 0.65 3.1.3. Debonding at concrete-matrix interface
s0 0.15 0.17 0.16 0.15
b 43.85 – 6.14 6.00
Fig. 9c shows the debonding failure mode at the concrete-
k(2) 0.73 0.74 – matrix interface. Loss of composite action by debonding of the
c – 1 – FRCM from the concrete substrate initiates at the region of maxi-
sr – 0.58 – mum bending moment after formation of flexural cracks or
flexure-shear cracks. As the load progresses, debonding extends
along the concrete-matrix interface towards the end of the
strengthening layer [11,23,25–29]. This could result in a clean
debonding of the FRCM composite from the concrete surface, or
debonding of the composite with a thin layer of concrete attached
to the composite [27].

3.1.4. Slippage of fabric roving within matrix


Slippage of the fabric usually precedes or accompanies debond-
ing at the fabric-matrix interface. The relative slip of fabric rovings
within the matrix increases with widening of cracks [27]. This may
explain why debonding typically initiates at the maximum
moment region. However, some specimens could fail solely by slip-
page of the fabric rovings, as reported by Loreto et al. [24], and
Babaeidarabad et al. [25] in the region of maximum bending
moment in specimens with one layer of FRCM, and by Ebead
et al. [28] in specimens with one or two layers of carbon-FRCM.
Fig. 8. Typical FRCM bond-slip relationship.
3.1.5. Rupture of fabric within matrix
This failure mode was observed in Elsanadeddy et al. [29], Sch-
tested by Jung et al. [21] and by other researchers [23–25]. The pre- laditz et al. [30], and Yin et al. [31]. The failure mode is shown in
dicted loads were in good agreement with those measured exper- Fig. 9d. Yin et al. [31] reported that with low FRCM reinforcement
imentally, having predicted-to-measured load ratios ranging ratio, concrete crushing could occur at the onset of rupture of the
between 0.85 and 1.20. fabric, without rupture of the tensile steel. However, with high
FRCM reinforcement ratio, rupture occurred in the fabric, and also
3. Flexural strengthening with FRCM in the tensile steel, without crushing of the concrete.

Table 3 gives a summary of previous studies [11,21,23–31] on 3.2. Factors affecting the flexural capacity of FRCM-strengthened
flexural strengthening of RC elements with FRCM composites. A members
discussion of the results of these studies is given in the following
sections. 3.2.1. Number of FRCM layers
Many studies have shown that increasing the number of FRCM
3.1. Failure modes layers resulted in an increase in the flexural capacity of strength-
ened specimens. The strength gain was usually non-proportional
Failure of RC elements strengthened in flexure with FRCM can to the number of FRCM layers. Ombres [23] recorded gains in the
occur due to concrete crushing after yielding of the tensile steel; load capacity of up to 16%, 33%, and 40%, for specimens with one,
shear failure due to the FRCM strengthening layer successfully two, and three layers, respectively. Schladitz et al. [30] reported
shifting the failure mode to shear; and rupture of the tensile steel increases in bearing load of one-way slabs corresponding to 67%,
in cases involving corrosion. The failure modes explained in the fol- 142%, 189%, and 245%, for specimens with one, two, three, and four
lowing sections are those pertaining to failure of the FRCM FRCM layers, respectively. Loreto et al. [24] recorded increases in
strengthening layers. the load capacity of up to 41% and 112%, for one and four layers
of FRCM respectively. El-Maaddawy and El Refai [26] recorded
3.1.1. Concrete cover separation increases in the load capacity of corroded beams of 39% and 51%,
This failure mode involves formation of an inclined crack at the for two and four layers respectively, measured with respect to that
end of the FRCM layer which extends horizontally along the level of a corroded unstrengthened beam. Yin et al. [31] observed that
of the tensile steel. With further increase in load, detachment of with one layer of FRCM, there was no improvement in fatigue life
the concrete cover occurs (Fig. 9a). El-Maaddawy and El Refai compared to that of a control specimen, however, with two and
[26] observed this failure mode in some specimens after the occur- three layers of FRCM, increases of 32% and 58%, respectively, were
rence of yielding of the tensile steel. recorded. The number of FRCM layers could affect the failure mode
of strengthened RC specimens. Ebead et al. [28] observed that spec-
3.1.2. Debonding at fabric-matrix interface imens with one or two layers of carbon-FRCM failed by slippage of
Failure of strengthened specimens by interfacial debonding at the fabric in the matrix, while specimens with three layers failed
the fabric-matrix interface initiates by the formation of a longitu- by debonding of the FRCM at the concrete-matric interface.
Table 3
Summary of previous studies on flexural strengthening with FRCM.

Reference
Hashemi Jung et al. Ombres Loreto et al. Babaeidarabad El-Maaddawy and D’Ambrisi and Ebead et al. [28] Elsanadedy Schladitz Yin et al. [31]
and Al- [21] [23] [24] et al. [25] El Refai [26] Focacci [27] et al. [29] et al. [30]
Mahaidi
[11]
No. of specimens 5 7 12 18 18 8 31 12 5 5 15
Test parameters Fabric Number of Number of Number of Number of Number of layers, Fabric type, number of layers, Steel reinforcement Number of Number of FRCM configuration,
type, layers, layers layers, layers fabric type, FRCM mortar type, FRCM ratio, fabric type, layers, layers number of layers, pre-
anchorage fabric type concrete configuration configuration, anchorage number of layers mortar type damage, corrosion
strength
Test method 4-Point 4-Point 4-Point 3-Point 3-Point 4-Point bending 4-Point bending, 3-point 4-Point bending 4-Point 4-Point 4-Point bending
bending bending bending bending bending bending bending bending
Loading protocol Monotonic Monotonic Monotonic Monotonic Cyclic Monotonic Monotonic Monotonic Monotonic Monotonic Monotonic
Concrete Cross-section type R-beam R-beam R-beam Slab R-beam T-beam R-beam R-beam R-beam Slab R-beam
bw  h (mm) 140  260 170  300 150  250 305  152 152  305 200  260 400  250 150  260 150  200 1000  230 120  240, 120  230
Length (mm) 7000 3000 3000 1830 1829 3200 N/A 2500 2200 700 2400
d (mm) 224 N/A 216, 217 129 260 200 N/A 210 164 N/A 193–195
Comp. strength 38 (cyl.) 28 (cyl.) 22–23 29, 43 (cyl.) 29, 43 (cyl.) 25 (cyl.) 48, 64 (cube) 67.5 (cyl.) 20 46 (cube) 45 (cube)
(MPa) (cube)

O. Awani et al. / Construction and Building Materials 132 (2017) 94–111


Tensile steel Bar type 3 Ø12 2 Ø10 3 Ø 12, 2 3 Ø9.5 (#3) 2 Ø13 3 Ø12 3 Ø14, 3 Ø16 2 Ø12, 2 Ø16 2 Ø10 Ø10 2 Ø12, Ø14, Ø16
Ø10
q (%) 1.08 N/A 1.05, 1.04 0.54 0.67 0.85 N/A 0.72, 1.27 0.64 N/A 0.98–1.72
fy (MPa) N/A 480 515, 526 414 276 538 476, 523 601, 595 372 574 545, 500, 415
Details of FRCM Fabric Fabric type Carbon Carbon PBO PBO PBO Carbon, basalt Carbon, PBO Carbon, PBO Basalt Carbon Carbon
+ glass
ff (MPa) N/A 4300 5800 N/A 276/751,3 3800, 1040 3051, 5213/53911, 4995/ 4300, 5800 623 1200 4100
50851
Ef (GPa) N/A 240 270 N/A N/A 230, 10 238–273 240, 270 32 N/A 180
tf (mm) 0.176 0.0107, 0.046/ 0.046/0.0112 0.046/0.0122 N/A 0.011–0.046 N/A 0.064 N/A N/A
0.0162 0.0222
bf (mm) 80⁄ 170 150 305 152 150 340 150 150 1000 120
efu (%) N/A 1.75 2.15 N/A N/A 1.6, 10.3 1.28–1.97 1.75, 2.15 N/A 12 2.3
Mortar fc,m (MPa) 65 (cube) 45 (cube) 29 N/A 34 (cube) 60 (cyl.) 16, 30+ 19.67, 29.06 24, 56 89+ 53 (cube)
(cube)
ft,m (MPa) 5.7* N/A 3.5 N/A N/A 2.1** 3, 5* N/A 2.8, 3.4*** 5.7* N/A
Em (GPa) N/A 40 6 N/A N/A 25 6.1, 7.5 N/A N/A N/A N/A
tm (mm) N/A 2 N/A N/A 5 N/A 6–8/layer 2–5 2 3 N/A
Composite FRCM EB EB EB EB EB EB, IE, EB+IE EB, EB-UW EB EB-UW EB EB, EB-UW
configuration
No. of layers 2 1–3 1–3 1, 4 1, 4 1, 2, 4 1–4 1–3 5, 10 1–4 1–3
ffrcm (MPa) N/A N/A N/A 1058 1664 N/A N/A 0.97, 1.55 7.7, 8.2 N/A N/A
Efrcm (GPa) N/A N/A N/A 137 128 N/A N/A 75, 121 N/A N/A N/A
efrcm (%) N/A N/A N/A 0.64 1.8 N/A N/A 1.25, 1.40 0.03,0.05 N/A N/A
tfrcm (mm) 20 5 N/A N/A N/A 6 N/A 10 N/A 6 10
bfrcm (mm) 140 170 150 305 152 150 340 150 150 1000 120
Height (mm) N/A N/A N/A N/A N/A N/A  250 N/A 125 N/A 110
Loading regime Effective span 2300 2700 2700 1524 1524 3000 2200, 1600 2200 2000 6750 2200
(mm)
Shear span (mm) 700 900 900 762 762 1300 750, 800 825 800 2750 700
Results Failure modes DB2, DB3 DB3, CS CC, DB1, FS, DB1, DB3 FS, DB3 CC, CS, DB2, DB3, SR DB2, DB3,S CC, FS, DB2, DB3 FR, DB3 FR CC, FR, SR
DB3
Strength gain (%) 9–28 32–120 9–44 35–112 13–92 7–93 1–54 8–78 39–91 67–245 0–58 (fatigue life)
Normalized N/A N/A 0.47–1.52 0.23–0.40 0.25–0.37 0.7–1.47 N/A 0.28–0.93 0.86–0.95 N/A N/A
ductility ratio

cyl. = cylinder compressive strength of concrete; cube = cube compressive strength; PBO = poly-p-phenylenebenzobisoxazole.
Failure modes: CC = concrete crushing after yielding of tensile steel; CS = concrete cover separation; DB1 = debonding of FRCM with thin layer of concrete; DB2 = debonding at fabric-matrix interface; DB3 = debonding at concrete-
matrix interface; FR = fabric rupture; FS = fabric slippage; S = shear; SR = tensile steel rupture; FRCM configuration: EB = externally-bonded; EB-UW = externally-bonded layers extended along lateral surface; IE = internally-
embedded; bfrcm = width of FRCM composite; bw = width of beam web; d = depth of tensile steel; Efrcm = Young’s modulus of FRCM composite; ffrcm = tensile strength of FRCM composite; fy = yield strength of longitudinal steel;
h = total section depth; tfrcm = total thickness of one layer of FRCM composite; efrcm = ultimate tensile strain of FRCM; q = tensile steel reinforcement ratio.
*
Determined from flexure test.
**
Determined from split cylinder test.
***
Determined from tensile testing of briquettes.
+
Determined using halved prisms from flexure test.
1
Tensile strength of fibers in longitudinal/transversal directions.
2
Nominal fiber bundle thickness in longitudinal/transversal directions.
3
Ultimate tensile strength per unit width of fabric in kN/m.

101
102 O. Awani et al. / Construction and Building Materials 132 (2017) 94–111

3.2.2. Fabric type load capacity gains of up to 28%. The use of fabric in a grid form
The type of fibers used in preparing the fabric has been found to (textile) with spacing between fiber tows allowed for better pene-
influence the flexural capacity of FRCM-strengthened members. tration of the matrix and hence better bond with the fabric, com-
D’Ambrisi and Focacci [27] compared the performance of carbon pared with those of continuous fabric sheets. The improved bond
fabric and PBO (poly-p-phenylenebenzobisoxazole) fabric in flexu- at the fabric matrix interface resulted in a higher strength gain.
ral strengthening of beams with FRCM. The former failed by slip-
page of the carbon fabric in the matrix, and the latter by 3.2.3. Mortar type
debonding at the concrete-matrix interface. The gains in flexural Proper selection of mortar to be used in FRCM strengthening is
capacity attained due to the carbon-FRCM and PBO-FRCM were crucial. The mix constituents of the mortar and overall mechanical
16% and 30% respectively, with respect to a control unstrengthened properties have significant effects on the performance of the FRCM
specimen. It was concluded that the PBO-FRCM performed better strengthening system. The inclusion of short fibers in the matrix
possibly due to a better bond between the PBO fabric and the could improve the performance of the mortar. Elsanadedy et al.
matrix. Ebead et al. [28] also observed greater ductility and energy [29] investigated the performance of an ordinary cementitious
absorption in the PBO-FRCM specimens compared to carbon-FRCM mortar and a polymer-modified cementitious mortar. The speci-
specimens. El-Maaddawy and El Refai [26], on the other hand, men with ordinary cementitious mortar failed by end-debonding
compared carbon fabric with basalt fabric in the strengthening of of the FRCM, while the specimen with polymer-modified mortar
corroded beams. The specimen with carbon-FRCM attained a load failed by rupture of the fabric. The gains in the load capacity of
increase of 51% while its counterpart basalt-FRCM attained a load the specimens were 82% and 91%, respectively, with respect to that
increase of only 7%, with respect to a corresponding unstrength- of a control unstrengthened specimen. It was concluded that the
ened corroded specimen. Thus the basalt fabric was less effective polymer-modified mortar provided better bond between the FRCM
than the carbon fabric in improving the beam capacity. This could layers and the concrete substrate, and hence, improved the beam
be attributed to the significantly lower tensile strength of basalt ductility, compared with those made of ordinary cementitious
fibers, compared to that of carbon fibers [32]. mortar.
Hashemi and Al-Mahaidi [11] strengthened two beams with
carbon sheets and two others with carbon textile. Results showed 3.2.4. FRCM configuration
that the beams strengthened with carbon sheets had load capacity In D’Ambrisi and Focacci [27], one beam was strengthened with
gains of up to 14%, relative to a control unstrengthened beam. On two layers of FRCM bonded at the soffit, while another beam was
the other hand, the beams strengthened with carbon textile had strengthened with a combination of one layer bonded at the soffit

Fig. 9. Typical flexural failure modes of FRCM-strengthened beams; (a) concrete cover separation, (b) debonding at fabric-matrix interface, (c) debonding at concrete-matrix
interface, and (d) rupture of fabric.
O. Awani et al. / Construction and Building Materials 132 (2017) 94–111 103

and one U-shaped layer. The specimen with two layers bonded at 43 MPa were used to fabricate the specimens. Results of the study
the soffit failed prematurely by debonding at the concrete-mortar indicated that varying the concrete strength had no significant
interface along with slippage of the fabric. The specimen with effect on the performance of the specimens. The researchers attrib-
one FRCM layer bonded at the bottom and one U-shaped FRCM uted this behaviour to a difference in the concrete covers of the
layer failed by slippage of the strengthening fabric. The strength specimens cast from the different concrete batches.
gains attained by the specimens were 9% and 18% respectively. This
may be an indication of greater effectiveness due to the inclusion 3.2.7. Corrosion
of U-shaped FRCM in flexural strengthening of RC beams. Yin Yin et al. [31] studied the effect of corrosion exposure on the
et al. [31] observed, however, that bonding at the soffit only was structural performance of carbon-FRCM-strengthened RC beams.
more effective than U-shaped bonding in enhancing the fatigue life The beams were subjected to wet-dry cycles for 6 months. In the
of beams. Further research is needed to fully understand the effect wet cycle, strengthened specimens were soaked in a 5% sodium
of U-shaped bonding on the effectiveness of FRCM to improve the chloride solution. Corrosion exposure reduced the fatigue life of
structural response and fatigue life of RC beams. strengthened beams. Corroded beams exhibited separation of the
In El-Maaddawy and El Refai [26], corroded T-beams were FRCM from the concrete substrate during testing. This phe-
repaired by first removing the deteriorated concrete and filling nomenon was not observed in non-corroded specimens indicating
the resulting groove with mortar to cover up the exposed steel. that corrosion weakened the bond between the FRCM layer and the
One beam was strengthened by applying two layers of carbon concrete substrate.
FRCM within the beam cover, while another was strengthened by
applying the FRCM layers externally. A third beam was strength- 3.2.8. Tensile steel reinforcement ratio
ened with four layers of FRCM applied internally within the cover, Ebead et al. [28] investigated the effect of tensile steel reinforce-
and a fourth with two layers applied internally and another two ment ratio (q) on the flexural performance of FRCM-strengthened
layers applied externally. The first two specimens failed by beams. Specimens with steel reinforcement ratios of 0.72% and
debonding at the fabric-matrix interface and at the concrete- 1.27% were tested. It was observed that the increase in load capac-
matrix interface, respectively. For the specimen with four FRCM ity, with respect to that of control unstrengthened specimens, was
layers internally-embedded within the concrete cover, failure higher in specimens with the lower reinforcement ratio. For a
was governed by cover separation. The fourth beam failed by a group of specimens, increases in load capacity corresponding to
combination of debonding at the fabric-matrix interface and cover 23–78% were associated with low reinforcement ratio, while
separation. The increase in the flexural capacity of the four speci- increases of 14–47% were associated with high reinforcement ratio.
mens, with respect to that of an unstrengthened corroded speci- The lower gain in capacity of specimens with high reinforcement
men, was 39%, 46%, 51% and 93%, respectively. All four ratio could be attributed to the greater contribution of the steel
strengthening schemes were effective, but the level of effective- reinforcement to flexural resistance.
ness varied depending on the FRCM configuration. The combina-
tion of both internally-embedded and externally-applied layers
was more effective than applying the same amount of FRCM only 4. Shear strengthening with FRCM
internally because of the increased bonded length of the external
layers, which delayed debonding and hence increased the flexural Table 4 gives a summary of previous studies [4,33–41] on shear
capacity. strengthening of RC elements with FRCM composites. A discussion
of the results of these studies is given in the following sections.
3.2.5. Anchorage
Anchoring of FRCM could help to improve their performance by 4.1. Failure modes
preventing or delaying the debonding mode of failure. Hashemi
and Al-Mahaidi [11] studied the effect of using two different 4.1.1. Cover separation
anchorage systems on the performance of FRCM-strengthened Separation of the concrete cover in shear testing of FRCM-
beams. The first anchorage system is described in Fig. 7. The second strengthened beams occurs by formation of longitudinal cracks
involved sandwiching the fabric between two steel plates. The on the top and/or bottom surface of the beams preceded by forma-
gripping action of the plates was achieved by means of two steel tion of several shear cracks in the shear span (Fig. 10). With further
bolts installed into concrete substrate. From the experimental increase in load, the longitudinal cracks widen and finally result in
study, it was observed that both anchored and unanchored separation of the lateral concrete covers of the beam [19,33,34]. In
strengthened beams failed by debonding at the fabric-matrix inter- Tetta et al. [34], this failure mode occurred in specimens with two
face, and there was no significant difference in the attained capac- to three layers of FRCM applied laterally or in U-shaped configura-
ities. This was attributed to the occurrence of debonding outside tion. The authors speculated that the possible offset of the fabric
the region of the anchorage system. layers when more than one fabric layer was used provided a denser
D’Ambrisi and Focacci [27] investigated the effect of using end mesh pattern, which led to a significantly improved mechanical
U-anchorage which consisted of U-shaped fabrics applied as an interlock, compared to the case of one fabric layer. However, spec-
extra layer at both ends of the FRCM strengthening system. Speci- imens with one or two layers failed by cover separation in Awani
mens with end U-anchorage failed by slippage of the carbon fabric et al. [33]. The occurrence of this failure mode, in this case, could
and attained an average gain in load capacity of 16%, relative to be due to a difference in stiffness between the FRCM reinforcement
that of a control unstrengthened beam. Specimens without anchor- and the concrete substrate, which led to a stress concentration and
age failed by a combination of debonding (concrete-matrix inter- high interfacial stresses between the lateral cover and the concrete
face) and slippage of the carbon fabric, with an average strength core.
gain of 9%.
4.1.2. Debonding at concrete-matrix interface
3.2.6. Concrete strength Debonding failure at concrete-matrix interface could occur by a
Loreto et al. [24] conducted flexural tests on one-way RC slabs clean separation of the strengthening layer from the concrete or
to investigate the effect of concrete strength on the effectiveness with a layer of concrete attached [35–37]. Specimens with high
of FRCM. Two batches of concrete with strengths of 29 MPa and FRCM reinforcement ratio (four to six layers) exhibited this failure
Table 4

104
Summary of previous studies on shear strengthening with FRCM.

Reference
Blanksvärd et al. [4] Awani Tetta et al. Brückner Contamine et al. Tzoura and Triantafillou and Al-Salloum Azam and Escrig
et al. [33] [34] et al. [35] [36] Triantafillou Papanicolaou et al. [39] Soudki [40] et al. [41]
[37] [38]
No. of specimens 21 9 9 12 10 11 4 10 7 9
Test parameters Mortar type, fabric Number of Number of Mechanical FRCM Fabric type, Angle of Angle of FRCM Fabric
type, concrete strength, layers, layers, FRCM anchorage, configuration, number of inclination of inclination of configuration, type
stirrup spacing, mortar stirrup configuration number of concrete layers, fabric, number of fabric, number of fabric type
thickness spacing layers strength, mechanical layers layers, mortar
application anchorage type
method
Test method 4-Point bending 3-Point 3-Point 3-Point 4-Point bending Lateral loading 4-Point bending 4-Point bending 3-Point 3-Point
bending bending bending bending bending
Loading protocol Monotonic Monotonic Monotonic Monotonic Monotonic Cyclic Monotonic, Monotonic Monotonic Monotonic
cyclic
Concrete Cross-section type R-beam R-beam R-beam T-beam R-beam T-beam R-beam R-beam R-beam R-beam
bw  h (mm) 180  500 150  300 102  203 120  450 150  250 150  350 150  300 150  200 150  350 300  300

O. Awani et al. / Construction and Building Materials 132 (2017) 94–111


Length (mm) 4500 4200 1677 2400 N/A 1000 2600 1500 2400 1700
d (mm) 418 250 177 372 220 320 272 164 308 254
Comp. strength (MPa) 49, 56, 74 (cube) 45 cube), 22 (cyl.) 27 (cube) 31, 42 (cube) 13, 23 (cube) 31 (cube) 20 (cyl.) 38 (cyl.) 34, 41
36 (cyl.) (cube)
Reinforcing Tensile Bar type 12 Ø16 6 Ø20 2 Ø16 6 Ø20 3 Ø20 3 Ø18 3 Ø16 4 Ø10 2 25M 3 Ø16
bars q (%) 3.21 5.04 2.23 4.22 2.86 1.59 1.48 1.28 2.2 0.79
fy (MPa) 555 520 547 500 570 545 575 578 480 517
Shear Bar type Ø12 Ø8 N/A Ø8 Ø6 N/A Ø5.5 N/A N/A N/A
Spacing (mm) 250, 350 75, 150 N/A 100, 200 200 N/A 230 N/A N/A N/A
qt (%) 0.50, 0.34 0.45, 0.22 N/A 0.84, 0.42 0.19 N/A 0.14 N/A N/A N/A
fyt (MPa) 601 294 N/A 500 570 N/A 275 N/A N/A N/A
Details of Fabric Fabric type Carbon Carbon Carbon AR-glass AR-glass Carbon Carbon Basalt Glass, carbon Basalt,
FRCM carbon,
PBO, glass
ff (MPa) 3800, 4300 3800 3800 556–590 N/A 3375 3350 623 2300, 3800 2610–
5800
Ef (GPa) 284, 288, 589 230 225 N/A N/A 225 225 32 75, 230 90–270
tf (mm) N/A N/A 0.095 N/A N/A 0.048, 0.096 0.047 0.064 N/A 0.042–
0.053
efu (%) N/A 1.6 N/A N/A N/A N/A N/A N/A 2.8, 1.6 1.8–3.2
Mortar fc,m (MPa) 22, 45, 77 74 (cube) 29+ 77+ N/A 22+ 31+ 24, 56 (cube) 58 (cube) 25, 30, 35+
ft,m (MPa) 5, 9, 11* 2.1** 9.4* 6.9* N/A 5* 4.2* 2.8, 3.4*** N/A 7.9, 8.6,
10.7*
Em (GPa) 18, 27, 35 22 N/A N/A N/A N/A 3.8 N/A N/A N/A
tm (mm) N/A N/A 2 2 N/A 2 1.5–2 2 N/A 4–5
Composite FRCM configuration SB SB SB, UW, FW UW SB, UW UW FW SB SB, UW UW
No. of layers 1 1, 2 1–3 2–6 N/A 1, 2 1,2 2, 4 1 1
ffrcm (MPa) N/A N/A N/A N/A 41 N/A N/A 7.7, 8.23 N/A N/A
Efrcm (GPa) N/A N/A N/A N/A 3 N/A N/A N/A N/A N/A
efrcm (%) N/A N/A N/A N/A 1.4 N/A N/A 0.03, 0.5 N/A N/A
tfrcm(mm) 20, 10 6 N/A N/A 1.7, 5 N/A N/A N/A 6–8 10
Height (mm) 500 300 203 330 250 250 N/A 200 350 285
Loading regime Effective span (mm) 4000 3000 1077 2000 2000 N/A 2200 1200 2000 1500
Shear span (mm) 1250 750 460 1000 700 800 775 400 1000 700, 800
Results Failure modes CC, FR CS CC, FS, CS S, DB3, CCS CC, DB1, DB3, FR FS, FR, DB3 S, F S S S
Strength gain (%) 42–184 51–145% 9–195 1–16 3–38 18–126 72–109 36–88 18–105 0–43

cyl. = cylinder compressive strength of concrete; cube = cube compressive strength; PBO = poly-p-phenylenebenzobisoxazole.
Failure modes: CC = concrete crushing after yielding of tensile steel; CCS = crushing of concrete strut; CS = concrete cover separation; DB1 = debonding of FRCM with thin layer of concrete; DB3 = debonding at concrete-matrix
interface; FR = fabric rupture; FS = fabric slippage; S = shear.
FRCM configuration: SB = bonded along lateral sides; UW = U-shape wrapped; FW = fully wrapped; fyt = yield strength of transverse steel (stirrups); qt = transverse steel (stirrup) reinforcement ratio.
*
Determined from flexure test.
**
Determined from split cylinder test
***
Determined from tensile testing of briquettes.
+
Determined using halved prisms from flexure test.
O. Awani et al. / Construction and Building Materials 132 (2017) 94–111 105

mode [35]. The premature debonding at concrete-matrix interface to two layers increased the strength gain by 32%, 5%, and 8%, for
was attributed to the high flexural rigidity of the FRCM jacket specimens without stirrups, specimens with stirrup spacing of
resulting in incompatible deformations with the concrete 150 mm, and specimens with stirrup spacing of 75 mm,
substrate. respectively.

4.1.3. Slippage of fabric roving within the matrix 4.2.2. Fabric type
Tzoura and Triantafillou [37] observed pulling/slippage of fabric Fabric geometry and type of fibers used have been shown to
rovings crossing shear cracks. This failure mode was accompanied influence the structural response of FRCM-shear-strengthened ele-
by partial rupture of fabric in the matrix, and occurred in speci- ments. Blanksvärd et al. [4] investigated the effect of fabric geom-
mens with one FRCM layer in Tetta et al. [34]. This failure mode etry using three carbon fabrics with different grid spacings and
was attributed to an insufficient interlock between the mortar mechanical properties. Using the fabric with the smaller grid spac-
and the fabric. When two to three layers were used, the interlock ing resulted in a higher first shear crack load. This was attributed to
was improved and failure was shifted to separation of the concrete the improved bond at the fabric-matrix interface, which led to for-
covers. mation of band of cracks in the shear span during the test. Tzoura
and Triantafillou [37] found that the use of higher amount of fibers
4.1.4. Rupture of fabric within the matrix in the fabric rovings increased the strength gain. In terms of type of
Rupture of fabric in the matrix occurs in fabric rovings crossing fibers used, Azam and Soudki [40] observed that carbon fabric was
shear cracks [4,36]. In Blanksvärd et al. [4], the strains in the fabric more effective than glass fabric in improving the shear capacity.
reached the ultimate strain of the fibers, which indicated a good This can be attributed to the higher stiffness of the carbon fabric
bond between the fabric and the mortar. The presence of mechan- compared with that of the glass fabric. Escrig et al. [41] compared
ical anchorage could also lead to rupture of fabric near the anchors, the performance of specimens shear-strengthened with glass-
as was observed in Tzoura and Triantafillou [37]. FRCM, basalt-FRCM, carbon-FRCM, and PBO-FRCM. Results of this
study showed that the PBO-FRCM was the most effective type of
fabric in increasing the shear capacity of the tested specimens.
4.2. Factors affecting the shear capacity of FRCM-strengthened
The carbon-FRCM exhibited inconsistent structural behaviour,
members
and this was attributed to a weak bond between the carbon fabric
and mortar. The effectiveness of the glass-FRCM and basalt-FRCM
4.2.1. Number of FRCM layers (reinforcement ratio)
in improving the shear capacity was comparable to that of the
A non-proportional increase in the shear capacity occurs as a
PBO-FRCM, despite the lower tensile strengths and Young’s moduli
result of increasing the number of layers. With two or more layers,
of the glass and basalt fabrics compared with those of the PBO fab-
a denser mesh pattern is created (particularly if the layers are off-
ric. This suggests that other factors (such as fabric-matrix bond)
set), leading to a better mechanical interlock in the FRCM system
influence the effectiveness of FRCM in shear strengthening.
and prevented the premature failure of the fabric [34]. Triantafillou
and Papanicolaou [38] reported strength gains of 72% and 109% for
one and two layers of FRCM, respectively, though both specimens 4.2.3. Mortar type
were tested differently (monotonic or cyclic loading). Specimens The incorporation of fibers in mortars or the use of polymer-
with two, three, four, and six layers of FRCM had shear capacity modified mortars enhanced the effectiveness of FRCM strengthen-
increases of 6%, 7%, 13%, and 16%, respectively, in Brückner et al. ing systems to improve the shear resistance of RC beams [4,39]. In
[35]. In Al-Salloum et al. [39], higher strength gains and deflections Al-Salloum et al. [39], the strength gains in specimens with
were observed in specimens with four layers of FRCM than those of polymer-modified mortar were up to 69% higher than those with
similar specimens with two layers. Specimens with two layers and regular cementitious mortar.
four layers experienced strength gains in the range of 36–37% and
46–88%, respectively. In Tzoura and Triantafillou [37], beams 4.2.4. FRCM configuration
strengthened with two layers of FRCM exhibited strength gains The conventional configuration of FRCM systems in shear
of up to 65%, relative to that of a control unstrengthened beam, strengthening of RC beams involves placing the rovings in a direc-
while those with one layer exhibited gains of up to 42%. In Tetta tion perpendicular orientation to the beam axis. A spiral applica-
et al. [34], comparing similar specimens, the strength gains caused tion can also be used, with the rovings placed at an angle to the
by one, two, and three layers of FRCM applied on the lateral sides beam axis. Triantafillou and Papanicolaou [38] concluded that
were 9%, 71%, and 110%, respectively, while the corresponding val- there was no major difference between the performance of the
ues for U-wrapped FRCM were 51%, 132%, and 153%. Awani et al. conventional and spiral configurations. However, Al-Salloum
[33] reported that increasing the number of FRCM from one layer et al. [39] reported that with higher number of FRCM layers, spiral

Front view Bottom view


Fig. 10. Concrete cover separation of an FRCM-strengthened beam (images taken by T. El-Maaddawy).
106 O. Awani et al. / Construction and Building Materials 132 (2017) 94–111

Steel bars anchored Anchor bolts at 45o


through flange through flange

FRCM FRCM

(a) (b)
Fig. 11. Typical mechanical anchorage systems of FRCM in T-beams; (a) bars anchored vertically through the flange, (b) anchor bolts inserted at 45° through the flange.

configuration resulted in strength gains of up to 1.5 times of those of FRCM systems to improve the structural response of RC
of the conventional configuration. U-shaped wrapping of fabric in elements.
FRCM shear strengthening could provide higher shear strength
gain than that provided by lateral application of the fabric [34]. 4.2.7. Stirrup spacing
In contrast, results from Azam and Soudki [40] showed similar per- Blanksvärd et al. [4] reported that with higher internal shear
formance of both configurations. Full-wrapping of the FRCM com- reinforcement ratio (smaller stirrup spacing), lower shear strength
posite is very effective and can successfully shift the failure mode gains may be achieved. This is in agreement with a more recent
from shear to flexural [34]. This configuration is, however, less eco- study conducted by Awani et al. [33]. Strength gains of 110%,
nomical and may not be feasible in some situations. 64%, and 51% were reported by Awani et al. [33], for specimens
without stirrups, specimens with stirrup spacing of 150 mm, and
4.2.5. Mechanical anchorage specimens with stirrup spacing of 75 mm, respectively. These val-
Fig. 11 shows typical mechanical anchorage systems for T- ues were for specimens with one FRCM layer. With two FRCM lay-
beams shear-strengthened with FRCM. Brückner et al. [35] ers, the corresponding values were 145%, 67%, and 55%. The shear-
strengthened T-beams with FRCM anchored with L-shaped steel strength gain is the ratio of the additional shear resistance pro-
sections located at the corner between the flange and the strength- vided by the FRCM reinforcement to the original shear resistance
ened web surface. The steel sections were glued to the FRCM sur- of the beam without strengthening. Specimens with a greater
face with epoxy adhesive, and anchored to the flange with amount of internal steel stirrups had higher original shear resis-
prestressed steel bars at 90° though the flange (Fig. 11a). The tance and, hence, exhibited a lower shear-strength gain.
anchorage system was found to increase strength gain and prevent
debonding mode of failure in specimens with higher numbers of 4.2.8. Mortar thickness
FRCM layers. Tzoura and Triantafillou [37] used an anchorage sys- Contamine et al. [36] reported that the thickness of FRCM layer
tem consisting of curved steel sections and 6 mm diameter did not significantly affect the shear strength gain of strengthened
threaded anchor bolts. The steel sections were located at the cor- specimens. Blanksvärd et al. [4] tested two FRCM shear-
ners between the flange and the web. The bolts were inserted in strengthened beams with a total mortar thickness of 10 mm and
epoxy-filled holes drilled at 45° and a depth of 80 mm in the flange, 20 mm, while keeping the amount of fabric unaltered. Both speci-
with spacings of 100 mm or 150 mm along the longitudinal axis of mens failed in flexure, thus the effect of varying mortar thickness
the beam. The unanchored specimens had strength gains of up to on the shear capacity could not be determined. More research is
65% and failed mostly by debonding. The anchored specimens with needed to investigate the effect of mortar thickness on the shear
100 mm and 150 mm anchor spacings had strength gains of up to strength gain.
126% and 104%, respectively. The anchored specimens failed
mostly by pull-out of the fabric rovings. Thus, the effectiveness 5. Column confinement with FRCM
of the FRCM system in improving the shear capacity was signifi-
cantly enhanced when mechanical anchors were present, with fur- Table 5 gives a summary of previous studies [42–50] on con-
ther enhancement provided when the anchor spacing was smaller. finement of concrete columns with FRCM composites. A discussion
The improved performance of the specimens with mechanical of the results of these studies is given in the following sections.
anchorage can be attributed to the change in the mode of failure.
The presence of mechanical anchorage prevented the premature
5.1. Failure modes
debonding mode of failure and allowed the beams to develop a
higher shear capacity.
5.1.1. Fabric rupture
This mode of failure could occur due to buckling of longitudinal
4.2.6. Concrete strength steel bars (if present) and dilation of the concrete [42–48]. Failure
Blanksvärd et al. [4] and Contamine et al. [36] had conflicting is initiated after the formation of vertical cracks which widen until
results on the effect of concrete strength on the shear strength gain rupture of the fabric occurs. Crushing of the concrete could occur
due to FRCM strengthening. Results from Blanksvärd et al. [4] sug- before failure, but the crushed concrete is held in place by the
gested that strength gain due to FRCM strengthening may be smal- FRCM jacket, till rupture of the jacket takes place [46,47,49]. In Tri-
ler with higher concrete strengths, while Contamine et al. [36] antafillou et al. [42], failure of FRCM-confined specimens was char-
recorded higher strength gain with higher concrete strength. The acterized by rupture in a limited number of fiber bundles, which
reported data was very limited. More research is need to investi- propagated slowly to other fiber bundles, resulting in a ductile
gate the effect of varying the concrete strength of the effectiveness mode of failure. A similar observation was reported by Bournas
Table 5
Summary of previous studies on confinement of concrete columns with FRCM.

Reference
Triantafillou Bournas et al. [43] Bournas et. Colajanni Colajanni et al. [46] Ombres [47] Ombres and Trapko [49] Trapko [50]
et al. [42] al. [44] et al. [45] Verre [48]
No. of specimens 24 11 7 22 30 20 8 15 15
Test parameters Mortar type, Number of layers, Design of Cross- Cross-section shape, Number of layers, Eccentricity, Temperature Eccentricity,
number of tie spacing longitudinal section number of layers, FRCM number of FRCM
layers steel, fabric shape, corner radius, loading configuration, layers configuration,
type number of protocol concrete strength number of layers
layers
Test method Uniaxial Uniaxial Lateral Uniaxial Uniaxial compression Uniaxial Uniaxial Uniaxial Uniaxial
compression compression, lateral loading with compression compression compression compression compression
loading with constant
constant axial load axial load
Loading protocol Monotonic Monotonic, cyclic Cyclic Monotonic Monotonic, cyclic Monotonic Monotonic Monotonic Monotonic
Concrete Type Cylindrical, Prismatic Prismatic Cylindrical, Cylindrical, prismatic Cylindrical Prismatic Cylindrical Prismatic

O. Awani et al. / Construction and Building Materials 132 (2017) 94–111


prismatic prismatic with corbels
Diameter/cross-section 150/ 200  200, 250  250 154, 200/ 200/200  200, N/A 200  200 113 200  200
dimensions (mm) 250  250 250  250 200  200 200  400
Length (mm) 300, 700 380, 1600 1600 335, 425 600 N/A 1200, 1300 300 1500
d (mm) N/A 166, 215 225 N/A N/A N/A N/A N/A N/A
Comp. strength (MPa) N/A 25 (cube) 27 (cube) 25 (cyl) 17 (cyl) 15, 29 (cyl) 24, 40 (cyl) 23 (cyl) 56 (cube), 49
(cyl)
Reinforcing Long. Bar type N/A 4 Ø12, 4 Ø14 4 Ø14 N/A N/A N/A 4 Ø8, 4 Ø10 N/A 4 Ø12
bars fy (MPa) N/A 563, 372 523 N/A N/A N/A 312, 436 N/A 500
Trans. Bar type N/A Ø8 Ø8 N/A N/A N/A Ø8 N/A Ø6
Spacing (mm) N/A 100, 200 200 N/A N/A N/A 120 N/A >200
fyt (MPa) N/A 351 351 N/A N/A N/A N/A N/A 220
Details of Fabric Fabric type Carbon Carbon Carbon, glass PBO PBO PBO PBO PBO PBO
FRCM ff (MPa) 3350 3800 3800, 1700 5800 5800 5800 5800 5800 5800
Ef (GPa) 225 225 225, 70 270 270 270 270
tf (mm) 0.047 0.095 0.095, 0.089 0.045/ 0.045/0.0122 0.046/0.0222 0.046/0.0222 0.046 0.046
0.0122
efu (%) N/A N/A N/A 2.2 2.2 2 2 N/A N/A
Mortar fc,m (MPa) 9, 31+ 22+ 21+ 16 (cyl.) 16 (cyl.) 30 (cube) 30 (cube) 29 29
ft,m (MPa) 3.3, 4.2* 6.8* 6.5* N/A N/A N/A N/A 3.5 3.5
Em (GPa) N/A N/A N/A N/A N/A 6.1 6.1 6 6
tm (mm) 2 2 2 3–4 3–4 3 4 N/A N/A
Composite FRCM configuration Spiral 0°/90° 0°/90° 0°/90° 0°/90° 0°/90°, spiral 0°/90° 0°/90° 0°/90°
No. of layers 2, 3, 4 4, 6 4 2, 3 2, 3 1 2, 3, 4 1,2 2 1,2,3
ffrcm (MPa) N/A N/A N/A N/A N/A N/A 1664 N/A N/A
Efrcm (GPa) N/A N/A N/A N/A N/A N/A 128 N/A N/A
efrcm (%) N/A N/A N/A N/A N/A N/A 1.76 N/A N/A
tfrcm (mm) N/A N/A N/A N/A N/A N/A N/A N/A N/A
Height (mm) 300, 700 430 430, 600 335, 425 335, 425 N/A 500, 900 300 1500
Results Failure modes DB2, FR FR, FR-DB2, FR-BB FR-BB, LS-PO FSR, FR FSR, FR FR-DB2, FR-DB-E FR-BB, CC DB2 DB2
Strength gain (%) 25–77 50–106, 5.51 7–26 11–51 11–51 5–220 20–39 77–123 0–24

cyl. = cylinder compressive strength of concrete; cube = cube compressive strength; PBO = poly-p-phenylenebenzobisoxazole
Failure modes: BB = bar buckling; CC = crushing of concrete; DB2 = debonding at fabric-matrix interface; E = failure at ends of specimen (top and bottom); FR = fabric rupture; FSR = relative slippage between fibers in a roving;
LS = longitudinal splitting along splice length of steel bars; PO = pull-out bond failure of spliced steel bars
*
From flexure test.
+
Determined using halved prisms from flexure test.
1
Average strength increase in push and pull directions, of specimen with lateral cyclic loading, with respect to an unstrengthened specimen.
2
Nominal fiber bundle thickness in longitudinal/transversal directions.

107
108 O. Awani et al. / Construction and Building Materials 132 (2017) 94–111

et al. [43]. The non-uniform distribution of forces in fiber bundles columns that were 200 mm in diameter strengthened with two
due to inadequate impregnation could lead to a relative slip and three FRCM layers exhibited strength gains of 28% and 37%,
between the outer and inner fibers in the fiber bundles. With more respectively [45]. The number of FRCM applied could influence
fabric layers (four instead of two), the failure was less ductile, the failure mode of the specimens, as observed by Ombres and
likely due to the fact that with a thicker jacket, there was better Verre [48]. Specimens with one layer of FRCM failed by rupture
stress distribution and more fiber bundles that were significantly of the fabric preceded by crushing of the concrete and buckling
stressed before failure [42]. FRCM systems with mortars having a of reinforcing bars, while specimens with two layers of FRCM failed
low tensile strength are more prone to fabric rupture mode of fail- by concrete crushing without damage to the FRCM jacket.
ure. Similarly, fabrics with low tensile strengths could promote
rupture of the jacket. In Bournas et al. [44], a specimen with
glass-FRCM jacket failed by rupture of fabric, whereas, the failure 5.2.2. Fabric type
of a similar specimen with carbon-FRCM jacket was not governed Bournas et al. [44] compared the effectiveness of carbon-FRCM
by fabric rupture. In Ombres [47] the fabric dilated and ruptured and glass-FRCM jackets in confining rectangular concrete columns.
at the top and bottom ends of spirally-wrapped specimens, and The effectiveness of both types was quite similar, in terms of
progressive damage of the mortar occurred. Fabric rupture resulted improving the axial capacity and deformation. However, the spec-
in a loss of confining action and eventual failure of the specimens. imen strengthened with glass-FRCM failed by rupture of the jacket
preceded by buckling of steel bars and dilation of the concrete,
5.1.2. Debonding while the carbon-FRCM jacket remained intact till the test ended
The debonding mode of failure in FRCM-confined concrete col- (due to the piston reaching its maximum stroke). The rupture of
umns could occur at the end of the concrete lap [42,43] or at the the glass fabric was most likely due to the lower tensile strength
fabric overlap region [47,49]. The debonding typically occurs at of the glass fibers.
the fabric-matrix interface. It is often accompanied or followed
by fabric rupture, or considerable slippage of the fabric rovings.
5.2.3. Mortar type
In Triantafillou et al. [42], two different mortar types were used
5.1.3. Fiber slippage within rovings
in preparing the FRCM jackets. The compressive and tensile
Fiber slippage with rovings mode of failure involves excessive
strengths were 8.6 MPa and 3.3 MPa for Mortar 1, and 30.6 MPa
relative slip between the outer and inner fibers in a roving. It is also
and 4.2 MPa for Mortar 2, respectively. Specimens with Mortar 1
known as telescopic failure [44,45]. It could result from non-
failed by debonding of the FRCM from the concrete at the fabric
uniform distribution of forces in the fabric due to poor impregna-
ends, while specimens with Mortar 2 failed by rupture of the fabric.
tion of the fibers. In Colajanni et al. [45], fiber slippage occurred
The change of mode of failure was attributed to changing proper-
within rovings after formation of cracks in the overlap region of
ties of the mortar used. The FRCM jacket made of Mortar 2 was
the fabric.
more effective than that made of Mortar 1 in improving the axial
load capacity of the columns.
5.1.4. Concrete crushing
Strengthened columns with a high amount of FRCM (two lay-
ers) could fail by concrete crushing, which may be accompanied 5.2.4. FRCM configuration
by buckling of the internal steel bars, without damage to the FRCM The conventional method of wrapping concrete columns with
jacket [48]. When the number of FRCM layers was reduced to one FRCM involves aligning the strong fiber direction of the fabric per-
layer, the specimens failed by fabric rupture. Removal of the pendicular (90° orientation) to the longitudinal axis of the column.
undamaged jacket after testing was necessary to evidence crushing A comparison between spiral wrapping and conventional wrapping
of the concrete core. carried out by Ombres [47] showed that spiral wrapping was less
effective than conventional wrapping in improving the axial capac-
5.2. Factors affecting the axial capacity of FRCM-strengthened ity of concrete columns. Furthermore, placing the fabric spirally at
members an angle of 45°, with respect to the longitudinal axis of the column,
was less effective than placing it at angle of 30°. The strength gains
5.2.1. Number of FRCM layers caused by two and three layers of FRCM with 90° orientation were
The gain in strength due to FRCM confinement of columns 61% and 93%, respectively. Placing the fabric at 30° reduced the
increases with an increase in the number of applied FRCM layers strength gains to 21% and 35%, respectively. The strength gains
[42,43,45,47,48,50]. In Triantafillou et al. [42], a group of speci- were further reduced to 15% and 22%, respectively, when the fabric
mens had two or three FRCM layers, while another group of spec- was placed at 45°. Trapko [50] wrapped square-section columns
imens had two or four FRCM layers. In the first group, strength using one layer of FRCM with the main/stronger fiber direction
gains in the range of 36–57% were reported for specimens with placed parallel to the longitudinal axis of the column, as well as
two FRCM layers, and 74–77% for specimens with three FRCM lay- another layer conventionally placed with the main fiber direction
ers, relative to an unconfined specimen. In the second group, a perpendicular to the longitudinal axis of the column. The results
specimen with two FRCM layers gained a 40% increase in strength, showed the ineffectiveness of placing the main fiber direction par-
while its counterpart with four FRCM layers gained a 51% increase allel to the longitudinal axis.
in strength, relative to an unconfined specimen. The ultimate strain
of the confined concrete also increased with an increase in the
number of FRCM layers. Similar observations were made by Bour- 5.2.5. Concrete strength
nas et al. [43] where strength gains in the range of 72–84% were Ombres [47] studied the effect of varying the concrete strength
recorded for specimens with four FRCM layers, and 84–106% for on the effectiveness of the confinement provided by FRCM. Results
specimens with six FRCM layers. Increasing the number of FRCM showed that for conventionally-wrapped specimens (90° orienta-
layers from two to three layers improved the strength gain of tion) the strength gain decreased with an increase in concrete
150 mm diameter concrete columns from 29% to 51% [45]. The strength. However, for spirally-wrapped specimens (30° and 45°
impact of increasing the number of FRCM layers was less orientation), there was a general increase in strength gain as the
pronounced for specimens with a greater section size. Concrete concrete strength increased.
O. Awani et al. / Construction and Building Materials 132 (2017) 94–111 109

5.2.6. Tie spacing tricity values of 16 mm and 32 mm, measured from the longitudi-
In Bournas et al. [43], varying the amount of ties in concrete col- nal axis of the column to the load point, were used. This
umns did not have a significant effect on the strength gain caused corresponded to eccentricity ratios of 0.08 and 0.16, respectively.
by FRCM jackets. The strength gain was calculated with respect to Concentrically-loaded specimens were also tested. The author
the strength of a similar unstrengthened column. With tie spacing studied the effectiveness of the FRCM jackets in terms of energy
of 200 mm, the strength gains for specimens with four and six of destruction (area under plot of load versus shortening distance
FRCM layers were 50% and 57%, respectively. The corresponding of the columns). The energy of destruction decreased with an
strength gains for specimens with tie spacing of 100 mm were increase in the eccentricity, with respect to that of similar
59% and 66%. This behaviour can be attributed to the significant unstrengthened specimens. For specimens with one layer of FRCM,
amount of FRCM used in strengthening of the tested specimens reductions of 26% and 33% were recorded with eccentricity ratios
which could have concealed the effect of varying the tie spacing of 0.08 and 0.16, respectively. The corresponding reductions for
on the strength gain. the specimens with two layers of FRCM were 19% and 37%. The
strength gain of confined specimens also decreased with an
5.2.7. Temperature increase in the eccentricity [48]. The failure mode of the tested
Trapko [49] studied the effect of high temperature exposure on specimens was not influenced by the loading eccentricity [48,50].
cylindrical columns confined with FRCM. Strengthened specimens
were exposed to temperatures of 40 °C, 60 °C and 80 °C. It was con- 5.2.12. Loading protocol
cluded that the strength reduction due to temperature exposure Colajanni et al. [46] investigated the effect of loading protocol
was in the range of 5–10%. In contrast, specimens confined with (monotonic or cyclic) on the effectiveness of FRCM in improving
epoxy-based carbon-FRP jackets experienced up to an average of the axial performance of concrete columns. The stress-strain
30% reduction in strength due to temperature exposure. curves of specimens tested monotonically were identical to the
envelope curves of corresponding specimens tested under cyclic
5.2.8. Section size/shape loading. However, in some specimens tested under cyclic loading,
FRCM jackets are more effective in confining circular-section there was more stress degradation in the post-peak phase com-
columns than prismatic columns. This is because the confinement pared to that of specimens tested monotonically. This was attribu-
pressure is uniform in circular sections, whereas it is non-uniform ted to extensive cracking in the mortar due to cyclic loading.
in prismatic sections. In Colajanni et al. [46], the confinement effect
of FRCM jackets on columns with circular, square, and rectangular
sections was studied. In prismatic specimens, the FRCM jacket was 6. Conclusions and recommendations
most effective for square-section specimens than for specimens
with rectangular section. It was observed that the strength gain 6.1. Conclusions
in circular-section specimens (Ø = 200 mm) was higher than that
of square-section specimens (200  200 mm). Colajanni et al. Results of previous studies reviewed in this paper have shown
[45] reported that the strengths of FRCM-strengthened square col- the practicality and effectiveness of FRCM composites in the
umns were on average 20% lower than those of their counterparts strengthening of RC structures. Based on the literature survey,
with circular cross-section. the following conclusions can be made:

5.2.9. Layout of longitudinal steel  FRCM strengthening systems have a great potential to improve
Bournas et al. [44] investigated the effect of using continuous the flexural, shear, and axial capacities of RC elements.
longitudinal steel versus lap-spliced longitudinal steel on FRCM  FRCM composites generally exhibit a good bond with concrete
confinement of columns. Also the effect of the splice length was substrate, and bond failure tends to occur at the fabric-matrix
studied using lengths of 20do and 40do (where do is the bar diam- interface. The fabric-matrix bond can be improved by modifica-
eter). Buckling of the longitudinal steel occurred in the specimen tions to the fabrics grid spacings and by proper mortar selec-
with continuous bars, but did not result in failure of the specimen tion. Certain types of fabrics tend to have a better bond with
as the compressive forces were effectively sustained by the sur- mortar, e.g., glass fabrics have a good bond with mortar, though
rounding confined concrete (the FRCM jacket remained intact at have much lower tensile strengths than PBO or carbon fabrics.
the end of the test due to the piston reaching its maximum stroke). Thus fabric selection is also crucial.
The specimen with the shorter splice length of 20do exhibited pull-  Properly designed anchorage systems could delay debonding
out bond failure of the bars after formation of longitudinal splitting failure at fabric-matrix or concrete-matrix interface, thereby
cracks along the splice length. Failure was suppressed in the spec- resulting in a higher strength gain.
imen with the longer splice length of 40do till the end of the test as  An increase in FRCM bond length results in an increase in capac-
the actuator’s stroke was exhausted before failure could occur. The ity up to the effective bond length (Le), beyond which no further
strength increases of the specimens with continuous bars and increase in capacity occurs. This length lies between 150 and
spliced bars were 13% and 26% (ith splice length of 20do), 330 mm for FRCM, based on previous studies.
respectively.  Increasing the number of layers of FRCM reinforcement results
in a non-proportional increase in the strength gain.
5.2.10. Corner radius  Typical failure modes of flexural and shear strengthened RC ele-
Colajanni et al. [46] observed that there was no significant influ- ments are concrete cover separation, debonding at fabric-
ence of the corner radius of prismatic specimens on the strength matrix interface, debonding at concrete-matrix interface, slip-
gain due to FRCM confinement. Nevertheless, for rectangular spec- page of fabric rovings, and fabric rupture. Debonding at fabric-
imens with four FRCM layers, an increase in the corner radius from matrix interface is more predominant in flexural-strengthened
15 mm to 30 mm resulted in a 70% increase in the ultimate strain. RC elements than in shear-strengthened RC elements.
 Failure of FRCM-strengthened concrete elements is often influ-
5.2.11. Load eccentricity enced by test parameters including mechanical properties of
Trapko [50] studied the effect of loading eccentricity on the the FRCM composite, strengthening configuration, presence of
effectiveness of FRCM in confining square-section columns. Eccen- anchorage, loading protocol. Temperature exposure reduced
110 O. Awani et al. / Construction and Building Materials 132 (2017) 94–111

the strength of FRCM-strengthened RC elements. The strength L – bond length


reduction due to temperature exposure was less significant Le – effective bond length
compared with that recorded for epoxy-based FRP systems. pf – perimeter of bond surface
 The most typical failure mode of FRCM-strengthened columns is n – number of fabric layers
rupture of the fabric due to the buckling of longitudinal steel s – bond slip (displacement of fabric in mortar)
reinforcement and dilation of the concrete. This failure mode sf – final bond slip when value of bond stress is zero
is usually ductile owing to the gradual propagation of rupture smax – relative displacement between fabric and matrix at
from a limited number of fiber bundles to other fiber bundles failure
in the FRM reinforcement. Other failure modes include debond- sr – parameter based on curve fitting of experimental data
ing of the fabric at the end of the FRCM lap, fabric slippage, and tf – nominal thickness of fabric
concrete crushing without damage to the jacket in over- tf,total – total nominal thickness of the fabric
strengthened RC columns. tm – thickness of one layer of mortar
 Eccentricity of loading in FRCM-strengthened concrete columns tfrcm – total thickness of one layer of FRCM composite
reduces the strength gain. a – curve-fitting parameter which is a function of the area
under the ascending branch of the s-s curve, the peak bond
6.2. Recommendations for future studies stress, and the slip at peak bond stress
ab – bond parameter based on stress in fabric at debonding
Although the reviewed studies have shed light on the bond b – parameter based on curve fitting of experimental data
behaviour, flexural performance, and shear performance of RC bf – parameter that depends on ratio of width of FRCM to that of
members strengthened with FRCM, further research on other the concrete section
parameters is still required for a better understanding of strength- bL – parameter that depends on actual and effective bond length
ening with FRCM. Only few studies have focused on developing efrcm – ultimate tensile strain of FRCM
FRCM bond-slip models. The available models are specific to the efu – ultimate tensile strain of fiber
tested types of FRCM and may not accurately describe other FRCM emax – fiber strain at failure
systems. More research is needed to develop bond-slip models that c – parameter based on curve fitting of experimental data
can be generalized for various FRCM systems. The typical failure q – tensile steel reinforcement ratio
mode of bond specimens was not prevalent in the flexure and qt – transverse steel (stirrup) reinforcement ratio
shear tests; hence more examination is necessary to understand rdeb – stress in FRCM at debonding
the different phenomena resulting in different failure modes. Fur- Ø – diameter of concrete member
ther research is required to fully understand the effect of U-shaped s – bond stress
bonding of FRCM on failure mode and structural response of RC s0 – initial finite value of bond shear stress
beams. Very limited information is available in the literature on
the effect of varying the strength of the concrete substrate on the
structural response of FRCM-strengthened RC elements. The viabil- Acknowledgement
ity of FRCM to upgrade the capacity of deep beams, pre-damaged
or corroded RC elements has received little attention. None of pre- The authors would like to express their gratitude to the UAEU
vious studies investigated the structural behaviour of continuous for financing this project under Grant No. 31N232.
RC elements strengthened with FRCM. More research is needed
to fill this gap. The effectiveness of using FRCM composites on References
uneven surfaces has not been studied. Other areas requiring fur-
ther research include durability, fire resistance, and fatigue perfor- [1] J.G. Teng, J.F. Chen, T.S. Scott, Debonding failures in FRP-strengthened RC
beams: failure modes, existing research and future challenges, in: Proceedings
mance of FRCM-strengthened RC elements. of Workshop on Composites in Construction. A Reality, Capri, Italy, 2001.
[2] O. Büyüköztürk, O. Gunes, E. Karaca, Progress on understanding debonding
problems in reinforced concrete and steel members strengthened using FRP
Notations composites, Constr. Build. Mater. 18 (1) (2004) 9–19.
[3] T. El Maaddawy, K. Soudki, Strengthening of reinforced concrete slabs with
mechanically-anchored unbonded FRP system, Constr. Build. Mater. 22 (4)
A – regressing parameter used for the relations between the (2008) 444–455.
strain of fabric and slip at loaded end of bond area [4] T. Blanksvärd, B. Täljsten, A. Carolin, Shear strengthening of concrete structures
b – width of RC member with the use of mineral-based composites, J. Compos. Constr. 13 (1) (2009) 25–
34.
bf – width of fabric [5] T. El Maaddawy, S. Sherif, FRP composites for shear strengthening of reinforced
bfrcm – width of FRCM composite concrete deep beams with openings, Compos. Struct. 89 (1) (2009) 60–69.
bw – web width of beam [6] T. El-Maaddawy, B. El-Ariss, Behavior of concrete beams with short shear span
and web opening strengthened in shear with CFRP composites, J. Compos.
d – effective depth of cross-section Constr. 16 (1) (2012) 47–59.
do – steel bar diameter [7] T. El-Maaddawy, Y. Chekfeh, Retrofitting of severely shear-damaged concrete
Ef – Young’s modulus of fiber t-beams using externally bonded composites and mechanical end anchorage, J.
Compos. Constr. 16 (6) (2012) 693–704.
Efrcm – Young’s modulus of FRCM composite
[8] T. El-Maaddawy, Y. Chekfeh, Shear strengthening of T-beams with corroded
Em – Young’s modulus of mortar stirrups using composites, ACI Struct. J. 110 (5) (2013) 779–789.
0
fc – concrete strength [9] ACI Committee 549, Guide to Design and Construction of Externally Bonded
Fabric-Reinforced Cementitious Matrix (FRCM) Systems for Repair and
fc,m – compressive strength of mortar
Strengthening Concrete and Masonry Structures (ACI 549.4R-13), American
ff – tensile strength of fiber Concrete Institute, MI, USA, 2013.
ffrcm – tensile strength of FRCM composite [10] R. Ortlepp, U. Hampel, M. Curbach, A new approach for evaluating bond
ft,m – tensile strength of mortar capacity of TRC strengthening, Cem. Concr. Compos. 28 (7) (2006) 589–597.
[11] S. Hashemi, R. Al-Mahaidi, Investigation of bond strength and flexural
fy – yield strength of longitudinal steel behaviour of FRP-strengthened reinforced concrete beams using cement-
fyt – yield strength of transversal steel (stirrups) based adhesives, Aust. J. Struct. Eng. 11 (2) (2010) 129–139.
h – height of RC member [12] A. D’Ambrisi, L. Feo, F. Focacci, Experimental analysis on bond between PBO-
FRCM strengthening materials and concrete, Compos. B 44 (1) (2013) 524–
k – bond effectiveness factor 532.
O. Awani et al. / Construction and Building Materials 132 (2017) 94–111 111

[13] L.H. Sneed, T. D’Antino, C. Carloni, Investigation of bond behavior of [32] J. Sim, C. Park, Characteristics of basalt fiber as a strengthening material for
polyparaphenylene benzobisoxazole fiber-reinforced cementitious matrix concrete structures, Compos. B 36 (6) (2005) 504–512.
composite-concrete interface, ACI Mater. J. 111 (5) (2014) 569–580. [33] O. Awani, T. El-Maaddawy, A. El Refai, Numerical simulation and experimental
[14] O. Awani, A. El Refai, T. El-Maaddawy, Bond characteristics of carbon fabric- testing of concrete beams strengthened in shear with fabric-reinforced
reinforced cementitious matrix in double shear tests, Constr. Build. Mater. 101 cementitious matrix, J. Compos. Constr. (2016) 04016056.
(2015) 39–49. [34] Z.C. Tetta, L.N. Koutas, D.A. Bournas, Textile-reinforced mortar (TRM) versus
[15] T. D’Antino, C. Carloni, L.H. Sneed, C. Pellegrino, Fatigue and post-fatigue fiber-reinforced polymers (FRP) in shear strengthening of concrete beams,
behavior of PBO FRCM-concrete joints, Int. J. Fatigue 81 (2015) 91–104. Compos. B 77 (2015) 338–348.
[16] L. Ombres, Analysis of the bond between fabric reinforced cementitious mortar [35] A. Brückner, R. Ortlepp, M. Curbach, Anchoring of shear strengthening for T-
(FRCM) strengthening systems and concrete, Compos. B 69 (2015) 418–426. beams made of textile reinforced concrete (TRC), Mater. Struct. 41 (2) (2008)
[17] L.H. Sneed, T. D’Antino, C. Carloni, C. Pellegrino, A comparison of the bond 407–418.
behavior of PBO-FRCM composites determined by double-lap and single-lap [36] R. Contamine, A.S. Larbi, P. Hamelin, Identifying the contributing mechanisms
shear tests, Cem. Concr. Compos. 64 (2015) 37–48. of textile reinforced concrete (TRC) in the case of shear repairing damaged and
[18] T. D’Antino, L.H. Sneed, C. Carloni, C. Pellegrino, Effect of the inherent reinforced concrete beams, Eng. Struct. 46 (2013) 447–458.
eccentricity in single-lap direct-shear tests of PBO FRCM-concrete joints, [37] E. Tzoura, T.C. Triantafillou, Shear strengthening of reinforced concrete T-
Compos. Struct. 142 (2016) 117–129. beams under cyclic loading with TRM or FRP jackets, Mater. Struct. 49 (1–2)
[19] O.M. Awani, Shear strengthening of reinforced concrete beams using textile- (2016) 17–28.
reinforced mortar (Master’s thesis), 2015. Retrieved from http://scholarworks. [38] T.C. Triantafillou, C.G. Papanicolaou, Shear strengthening of reinforced
uaeu.ac.ae/all_theses/213. concrete members with textile reinforced mortar (TRM) jackets, Mater.
[20] A. D’Ambrisi, L. Feo, F. Focacci, Bond-slip relations for PBO-FRCM materials Struct. 39 (1) (2006) 93–103.
externally bonded to concrete, Compos. B 43 (8) (2012) 2938–2949. [39] Y.A. Al-Salloum, H.M. Elsanadedy, S.H. Alsayed, R.A. Iqbal, Experimental and
[21] K. Jung, K. Hong, S. Han, J. Park, J. Kim, Prediction of flexural capacity of RC numerical study for the shear strengthening of reinforced concrete beams
beams strengthened in flexure with FRP fabric and cementitious matrix, Int. J. using textile-reinforced mortar, J. Compos. Constr. 16 (1) (2012) 74–90.
Polym. Sci. 2015 (2015) 868541. [40] R. Azam, K. Soudki, FRCM strengthening of shear-critical RC beams, J. Compos.
[22] J.G. Teng, S.T. Smith, J. Yao, J.F. Chen, Intermediate crack-induced debonding in Constr. 18 (5) (2014) 04014012.
RC beams and slabs, Constr. Build. Mater. 17 (6) (2003) 447–462. [41] C. Escrig, L. Gil, E. Bernat-Maso, F. Puigvert, Experimental and analytical study
[23] L. Ombres, Flexural analysis of reinforced concrete beams strengthened with a of reinforced concrete beams shear strengthened with different types of
cement based high strength composite material, Compos. Struct. 94 (1) (2011) textile-reinforced mortar, Constr. Build. Mater. 83 (2015) 248–260.
143–155. [42] T.C. Triantafillou, C.G. Papanicolaou, P. Zissimopoulos, T. Laourdekis, Concrete
[24] G. Loreto, L. Leardini, D. Arboleda, A. Nanni, Performance of RC slab-type confinement with textile-reinforced mortar jackets, ACI Struct. J. 103 (1)
elements strengthened with fabric-reinforced cementitious-matrix (2006) 28–37.
composites, J. Compos. Constr. 18 (3) (2014) A4013003. [43] D.A. Bournas, P.V. Lontou, C.G. Papanicolaou, T.C. Triantafillou, Textile-
[25] S. Babaeidarabad, G. Loreto, A. Nanni, Flexural strengthening of RC beams with reinforced mortar versus fiber-reinforced polymer confinement in reinforced
an externally bonded fabric-reinforced cementitious matrix, J. Compos. Constr. concrete columns, ACI Struct. J. 104 (6) (2007) 740–748.
18 (5) (2014) 04014009. [44] D.A. Bournas, T.C. Triantafillou, K. Zygouris, F. Stavropoulos, Textile-reinforced
[26] T. El-Maaddawy, A. El Refai, Innovative repair of severely corroded T-beams mortar versus FRP jacketing in seismic retrofitting of RC columns with
using fabric-reinforced cementitious matrix, J. Compos. Constr. 20 (3) (2016) continuous or lap-spliced deformed bars, J. Compos. Constr. 13 (5) (2009) 360–
04015073. 371.
[27] A. D’Ambrisi, F. Focacci, Flexural strengthening of RC beams with cement- [45] P. Colajanni, F. De Domenico, A. Recupero, N. Spinella, Concrete columns
based composites, J. Compos. Constr. 15 (5) (2011) 707–720. confined with fibre reinforced cementitious mortars: experimentation and
[28] U. Ebead, K.C. Shrestha, M.S. Afzal, A. El Refai, A. Nanni, Effectiveness of fabric- modelling, Constr. Build. Mater. 52 (2014) 375–384.
reinforced cementitious matrix in strengthening reinforced concrete beams, J. [46] P. Colajanni, M. Fossetti, G. Macaluso, Effects of confinement level, cross-
Compos. Constr. (2016) 04016084. section shape and corner radius on the cyclic behavior of CFRCM confined
[29] H.M. Elsanadedy, T.H. Almusallam, S.H. Alsayed, Y.A. Al-Salloum, Flexural concrete columns, Constr. Build. Mater. 55 (2014) 379–389.
strengthening of RC beams using textile reinforced mortar–experimental and [47] L. Ombres, Concrete confinement with a cement based high strength
numerical study, Compos. Struct. 97 (2013) 40–55. composite material, Compos. Struct. 109 (2014) 294–304.
[30] F. Schladitz, M. Frenzel, D. Ehlig, M. Curbach, Bending load capacity of [48] L. Ombres, S. Verre, Structural behaviour of fabric reinforced cementitious
reinforced concrete slabs strengthened with textile reinforced concrete, Eng. matrix (FRCM) strengthened concrete columns under eccentric loading,
Struct. 40 (2012) 317–326. Compos. B 75 (2015) 235–249.
[31] S.P. Yin, J. Sheng, X.X. Wang, S.G. Li, Experimental investigations of the bending [49] T. Trapko, The effect of high temperature on the performance of CFRP and
fatigue performance of TRC-strengthened RC beams in conventional FRCM confined concrete elements, Compos. B 54 (2013) 138–145.
and aggressive chlorate environments, J. Compos. Constr. 20 (2) (2016) [50] T. Trapko, Effect of eccentric compression loading on the strains of FRCM
04015051. confined concrete columns, Constr. Build. Mater. 61 (2014) 97–105.

You might also like