Sunteți pe pagina 1din 10

EFFECT OF CONCRETE CRACK WIDTH ON CORROSION OF EMBEDDED REINFORCEMENT William H. Hartt, PhD, PE Hartt and Associates, Inc.

20914 Morada Court Boca Raton, Florida 33433 March 18, 2009 BACKGROUND It is generally recognized that concrete cracks above a certain width facilitate chloride ingress and initiation of active corrosion of embedded reinforcement sooner than would otherwise occur in sound concrete. This has led professional societies and certification authorities to establish design criteria such that width of cracks on the exterior concrete surface does not exceed a specified value. For a sea water splash zone exposure, multiple authors have recommended that this maximum crack width be 0.3 mm.1,2,3 At the same time, the correlation between crack width and extent of corrosion is not particularly strong.4,5 If it can be demonstrated that reinforcement corrosion in hostile exposures is not strongly sensitive to crack width, then a case can be made for relaxing current maximum crack width criteria to 0.45 mm provided corrosion resistant reinforcement such as MMFX2 is specified. The purpose of this commentary is to review the existing literature pertaining to the effect of concrete crack width on reinforcement corrosion in environments such as exist in the Middle East and project appropriateness of such a relaxation. Both marine and inland structures invariable encounter different exposure zones. In the former case (marine), these are submerged, tidal, splash, and atmospheric. For the latter (inland), zones are categorized as 1) buried below the water table, 2) buried in the water table transition zone (depth range where the water table level cycles), 3) buried in aerated soil, and 4) atmospheric (above ground level). For situations where the soil and ground water contain significant concentrations of chloride and sulfates, as is likely to be the case in many Middle East locations, similarities exist between zones of the two types of exposure (marine and inland); however, focus in the present commentary is placed upon structures that are located inland. Figure 1 provides a schematic illustration of an inland located concrete element that extends through all four of the above zones. The propensity for reinforcing steel corrosion at concrete cracks in each of these four zones is described below. EFFECT OF CORROSION ZONE Zone 1 Here, ground water with chlorides and sulfates rapidly penetrate to the full depth of cracks and potentially exceed the threshold concentration required to depassivate the reinforcing steel and initiate active corrosion. Rodriguez and Hooton6 reported, based upon experiments involving concrete specimens with artificially created cracks submerged in chloride solutions that chloride uptake was independent of crack width in the range investigated (0.08 to 0.68 mm) and was uniform with crack depth. Irrespective of chloride and sulfate concentration, however, corrosion rate is nil in this zone because dissolved oxygen concentration, a necessary requisite for corrosion, is zero in ground waters. Hence, concrete cracks are not a concern here.

Page 1 of 10

Zone 4 (Atmospheric) Ground Level Zone 3 (Aerated Soil)

Zone 2

Zone 1 (Ground Water)

Figure 1: Schematic illustration of a reinforced concrete element that transitions traversing different inland corrosion zones. Zone 2 In this zone, the concrete cycles between being wet and saturated during periods when the water table is high and relatively dry when the water table is low. Consequently, chlorides readily accumulate within concrete cracks as for Zone 1 during the former period (concrete is wet and saturated) and dissolved oxygen is readily available during times when the water table is low. Even when the water table is high, corrosion may occur because of a macro-cell between the reinforcement at the crack tip (anode) and more elevated reinforcement in dryer, oxygenated concrete (cathode). The extent of corrosion here is likely to be controlled by, first, the range of water table variability and, second, concrete quality and cover to the extent that these restrict macro-cell activity. Also, this corrosion may be limited somewhat by cathodic polarization via the more negative potential of reinforcement in Zone 1.7 Zone 3 Enhanced corrosion in this zone requires that chlorides and sulfates from the soil diffuse into the concrete along the crack faces. This is likely to most pronounced during periods when the soil is moist from rain. Consequently, the risk of corrosion is less here than for Zone 2. Zone 4 In this zone, chlorides and sulfates are present on the exterior concrete surface and to a lesser extent within cracks as a consequence of contact by airborne particulates. The detrimental ions can then diffuse into cracks and potentially initiate corrosion. CORROSION MECHANISM AT CONCRETE CRACKS The generally accepted mechanism whereby reinforcement corrosion occurs at concrete cracks involves a macro-cell where steel exposed at the crack tip and immediately thereto becomes an anode and actively corrodes once the critical chloride concentration is exceeded here with adjacent passive rebar serving as a cathode for oxygen reduction. Figure 2 illustrates this process schematically. In order to maintain charge conservation, the anodic corrosion reaction can occur only as rapidly as oxygen is
Page 2 of 10

reduced at cathodic sites. Consequently, good concrete quality (high resistivity) and greater concrete cover have been identified as important in limiting the extent of this macro-cell activity.8,9
Concrete Crack

Concrete

Rebar (2)

Figure 2: Schematic illustration of macro-cell corrosion activity at a concrete crack. SUMMARY OF RELEVANT LITERATURE Schiessl and Raupach These authors10 performed corrosion studies on mechanically cracked concrete specimens with crack widths in the range 0.1 0.5 mm. They observed that, while corrosion initiated earlier the wider the crack, in all cases macro-cell current activity generated by reinforcement corrosion at the crack decreased with time and in the long-term (two years) was independent of crack width. They concluded that concrete cover and quality (mix design) were much more important in controlling corrosion at cracks than was crack width. Debate: Crack Width, Cover, and Corrosion11 Concern regarding appropriateness of the ACI Building Code applicable in 1985, as this pertained to the interrelationship between concrete design, cracks, and corrosion, prompted a formal discussion by recognized authorities (D. Darwin, D.G. Manning, E. Hognestad, A.W. Beeby, P.F. Rice, and A.Q. Ghowrwal) which was subsequently published as indicated above. In this, Hognestad pointed out that results from Corps of Engineer experiments involving loaded reinforced concrete beams exposed at Treat Island for 25 years failed to indicate any correlation between crack width and extent of deterioration. This led him to conclude, the Treat Island tests warn us not to get over-excited about crack widths and steel stresses. This conclusion agrees with that of Schiessl and Raupach discussed above. Sags, Kranc, Presuel-Moreno, Rey, Torres-Acosta, and Yao12 These authors conducted a comprehensive investigation of corrosion on eight different coastal (marine exposed) reinforced concrete bridges in Florida. As a part of this study concrete cracks were identified and pairs of cores were acquired where one was taken centered on a crack and a second from sound concrete to the side of the crack at the same elevation. Some cores were from the splash and the remainder from the atmospheric zone. Crack widths were measured prior to coring, and powdered concrete samples were taken from each core pair at a common depth into the core from what had been the exposed surface and analyzed for acid soluble chlorides using the standardized Florida Department of Transportation method.13 The samples from the cracked cores were from along the crack tip. It was assumed that chloride concentration in the vicinity of the crack tip represents a measure of corrosion propensity for the reinforcement here. Figure 3 plots the results as chloride concentration ([Cl-])
Page 3 of 10

determined from sound concrete divided by [Cl-] at the same depth at the crack ([Cl-] (Sound))/([Cl-] (Crack)) in units of pounds per cubic yard (pcy) versus crack width. The data are broadly scattered with no indication that chloride concentration at the crack depended upon crack width.

100.0

HF Bridge Dames Point Bridge

CC Causeway Bay Side Bridge New Smyrna Bridge BC Bay Bridge

[Cl-] (Solid)/[Cl-] (Crack)

Safety Harbor Bridge New Pass Bridge

10.0

1.0

0.1 0 0.1 0.2 0.3 Crack Width, mm 0.4 0.5

Figure 3: Data from Florida bridges showing a lack of correlation between chloride concentration within concrete cracks and crack width. Berke, Dellaire, Hicks, and Hoopes14 In this study, the authors performed experiments on four cracked CEN (European Committee for Standardization) reinforced concrete specimens but with a crack being introduced by loading in three point bending. The CEN specimen uses smooth bars (no deformations), and so some slippage of the bars during loading is expected such that the crack extends to a greater depth than would be the case for deformed bars. While still loaded, a shim was positioned in the crack to minimize closure. The specimens were ponded with a 3 wt% NaCl solution on a two week wet two week dry cycle. They were dissected after 16 months and both the corroded area on the top bar and chloride concentration in the concrete at the top bar near the crack were measured. Figure 4 shows a plot of these data and reveals a crack width for three of the specimens as 0.25 mm (0.01 in.) and for the fourth 0.51 mm. (0.02 in.). Obviously, data scatter for correlating corroded area with crack width is large, indicating the variability that can be encountered here. Thus, for two of the smaller crack specimens, the corroded area was relatively small, whereas for the third, the corroded area was about the same as for the larger crack width specimen. Particularly noteworthy, however, is that chloride concentration at the crack tip region, which is the major factor in promoting depassivation of the reinforcement, was essentially the same for the two crack widths.

Page 4 of 10

9 10,000
Corroded Area, mm2

8 7
Corroded Area

8,000 6,000 4,000 2,000 0 0 0.1 0.2 0.3 0.4 Crack Width, mm 0.5 0.6
[Cl-]

6 5 4 3 2 1 0

Chloride Conc. At Crack Tip, kg/m3

12,000

10

Figure 4:

Correlation of the extent of corrosion and chloride concentration with crack width as reported by Berke et al.

PERFORMANCE OF MMFX2 COMPARED TO BLACK BAR IN CRACKED CONCRETE Presuel-Moreno and Roarke15 Results from a research study by these authors involved G109 type specimens fabricated with a simulated concrete crack of a specific width and reinforced with either BB and galvanized bar (GV). No MMFX2 reinforcement was employed in this study. The simulated crack was introduced by placing a 1.6 mm (0.063 in.) thick stainless steel sheet in the form to depth of the upper bar prior to concrete placement. The sheet was removed after initial concrete set. As such, crack width (1.6 mm) was constant along the entire span of the reinforcement cover. The specimens were ponded on a two week wet two week dry cycle with 15 wt% NaCl, which is considered a very harsh exposure. Figure 5 shows a plot of macro-cell current (proportional to corrosion rate) versus time for three concrete slab specimens reinforced with BB and three with GV. A sustained current of 10 A was taken as indicative of corrosion initiation. The data show that for both the BB and GV reinforcements there was a tendency for repassivation subsequent to activation, consistent with the reporting by others that corrosion at cracks often arrests after a period of time, apparently because of corrosion products sealing off the crack mouth.10,16 This makes definition of time-to-corrosion, Ti, difficult; however, Table 1 lists the projected Ti values for these specimens along with the ratio of Ti for GV compared to that for BB [Ti(GV)/Ti(BB)] based on the three specimen average Ti for the two reinforcement types, where the latter parameter assumes that the GV reinforced specimens activated at the indicated Ti. From experiments on reinforced concrete specimens exposed to chlorides and data collected from cracked locations on concrete bridge decks subjected to deicing salts in Kansas, Darwin et al.17 projected that corrosion of BB reinforcement with 76 mm cover initiates on average after 2.3 years, with galvanized rebar (GV) after 4.8 years, and with MMFX2 14.8 years. The corresponding Ti ratios relative to BB are Ti(GV)/Ti(BB) = 2.1 and Ti(MMFX2)/Ti(BB) = 6.4. No crack widths were reported in this study;
Page 5 of 10

however, the Ti(GV)/Ti(BB) value reported above by Presuel-Moreno and Roarke (4.9) is generally consistent with the one determined by Darwin et al.15 (2.1), considering differences in the two

30 25 20 15 10 5 0 0

CT41-GV-CRK1 CT41-GV-CRK2 CT41-GV-CRK3 CT41-BB-CRK1 CT41-BB-CRK2 CT41-BB-CRK3

Current, microA

500

1000

1500 2000 2500 Exposure Time, days

3000

3500

Figure 5:

Macro-cell current (corrosion rate) versus time for cracked concrete specimens reinforced with BB and GV. Table 1: Listing of Ti and average Ti(GV)/Ti(BB) for the data in Figure 6.
Specimen CT41-BB-CRK1 CT41-BB-CRK2 CT41-BB-CRK3 CT41-GV-CRK1 CT41-GV-CRK2 CT41-GV-CRK3 Ti 756 657 290 2,765 >2,765 >2,035 Ti(GV)/Ti(BB)

4.9

experimental approaches. On this basis, the projection by Darwin et al. of Ti(MMFX2)/Ti(BB) being 6.4 (see above) seems appropriate, even considering the relatively wide crack width involved (1.6 mm). Hartt, Powers, Lysogorski, Liroux, and Virmani18 The above analyses aside, it can be reasoned that the most conservative approach to evaluating the effect of crack width on durability of reinforced concrete is to consider that Ti is zero (corrosion initiates upon initial exposure) and service life is determined by the time for corrosion to propagate to the point where repair or rehabilitation is required. Results of a study by the above authors allow for a comparison of corrosion rates at simulated cracks for concrete specimens reinforced with MMFX2 compared to black bar (BB). In this, specimens were exposed outdoors in south Florida approximately 300 m inland from the Atlantic Ocean while ponded with a 15wt% NaCl solution on a one week wet one week dry cycle.
Page 6 of 10

Figure 6 provides a schematic illustration of the general specimen geometry, and Figure 7 shows a photograph of two specimens under test. For three BB and three MMFX2 specimens, a simulated concrete crack that intersected the top three bars was part of the fabrication. The cracks were produced the same as described above for the experiments of Presuel-Moreno and Roarke and were also 1.6 mm (0.063 in.) wide. Thus, as above, crack width was constant along the reinforcement cover. Macro-cell current between the three top and three bottom bars, which is directly proportional to corrosion rate of the top bars, was measured weekly. Crack Front NaCl Pond 30 15 Simulated Crack

All dimensions in cm. 30 Figure 6: Schematic illustration of the reinforced concrete specimens.

Figure 7: Photograph of two slab specimens with ponding baths under test. Figure 8 shows a plot of macro-cell current density versus time for the three BB and three MMFX2 concrete cracked specimens. The saw-toothed nature of the data resulted because acquisitions were alternately made when specimens were wet and then dry (high values correspond to the slabs being wet). Particularly noteworthy is that current density for the BB specimens tended to stabilize (become constant) after about 200 days, whereas for MMFX2 current density was lower than for BB and decreased with time beyond about 200 days and stabilized at a relatively low value after about 400 days. This difference is attributed to formation of a more dense corrosion product film within the crack on the MMFX2 which provided greater corrosion protection than for BB. Beyond 400 days, the average current density for the three BB specimens was 0.945 A/cm2, where for the three MMFX2 specimens the average current density was 0.0165 A/cm2. Consequently, the BB specimens were corroding 5.7 times more rapidly than the MMFX2 ones.
Page 7 of 10

1.75 1.50

Current Density, A/cm2

1.25 1.00 0.75 0.50 0.25 0.00 0 100

3-CCON-BB-1 3-CCON-BB-2 3-CCON-BB-3 3-CCON-MMFX-1 3-CCON-MMFX-2 3-CCON-MMFX-3

200

300

400

500

600

Exposure Time, days

Figure 8: Macro-cell current density data for three specimens each of BB and MMFX2. Subsequent to these exposures, one specimen with each reinforcement was dissected and the reinforcement examined for corrosion. Figure 9 shows the reinforcement and adjacent concrete for specimen 3-CCON-BB-1 and Figure 10 does the same for 3-CCON-MMFX-1. The extent of corrosion

Figure 9: Photograph of reinforcement from the three BB reinforced specimens. product staining is generally consistent with conclusions reached from the macro-cell current data (Figure 8); that is, corrosion was much less developed for the MMFX2 reinforcement compared to BB. Conclusions While intuition suggests that corrosion of reinforcement at the base of concrete cracks should increase with crack width, there is evidence in the literature that crack width does not significantly influence long-term durability of reinforced concrete. From the available information, as reviewed above, a conservative estimate is that MMFX2 reinforcement initiates corrosion at the base of concrete cracks at
Page 8 of 10

Figure 10: Photograph of reinforcement from the three MMFX2 reinforced specimens. exposure times about six times greater than for BB. This is likely to be the case for cracks as wide as 1.6 mm (0.063 in.). Also, corrosion rate in cracked concrete, once initiated, is about six times less for MMFX2 compared to BB. This also is projected to apply for crack widths as large as 1.6 mm (0.063 in.). Figure 11 provides a schematic illustration of this difference assuming that BB initiates corrosion at time 3 (arbitrary scale) and MMFX2 at time 18 (six times longer). It is concluded from this that it is appropriate to relax existing maximum crack width criteria to 0.45 mm provided corrosion resistant reinforcement such as MMFX2 is specified.

18
Cumulative Corrosion Danage, arbitrary units

16 14 12 10 8 6 4 2 0 0 20 40 60 80 Time, arbitrary units 100 120


BB MMFX2

Figure 11: Schematic illustration of cumulative corrosion damage to BB and MMFX 2 reinforcement at concrete cracks assuming a six-fold greater time-to-corrosion and a six-fold reduction in corrosion rate, once initiated, for MMFX2.

Page 9 of 10

References 1 R.D. Browne and A.F. Baker, The Performance of Structural Concrete in a Marine Environment, in Development of Concrete Technology, Ed. F.D. Lyden, Applied Science Pub., London, 1979, p. 111. 2 R.D. Browne, Low Maintenance Concrete Specification versus Practice?, in Proc. 2nd Intl. Conf. on Maintenance of Marine Structures, Thomas Telford, UK, 1986. 3 A.L. Marshall, Marine Concrete, Van Nostrand Reinhold, NY, 1990. 4 A. Bentur, S. Diamond, and N.S. Berke, Steel Corrosion in Concrete, E&FN Spon, London, 1997. 5 H. Martin and P. Schiessel, The Influence of Cracks on the Corrosion of Steel in Concrete, Intl. Symp. on Durability of Concrete, Rilem, 1969, p. D205. 6 O.G. Rodriguez and R.D. Hooten, Influence of Cracks on Chloride Ingres into Concrete, ACI Materials Journal, March-April, 2003, p. 120. 7 R.E. Tanner and W.H. Hartt, Specimen Type and Test Protocol for Corrosion Performance Characterization of Reinforced Concrete Marine Pilings, paper no. 08320, CORROSION/08, NACE International, Houston, 2008. 8 E.A. Baker, K.L. Money, and L.R. Sanborn, Marine Corrosion Behavior of Bare and Metallic-Coated Steel Reinforcing Rods in Concrete, in Chloride Corrosion of Steel in Concrete, STP 629, American Society for Testing and Materials, Phila., 1977, p. 30. 9 F.M. Lea and C.M. Watkins, The Durability of Reinforced Concrete in Sea Water, Research Paper No. 30, National Building Studies, Dept. of Scientific and Industrial Research, Her Majestys Stationery Office, London, 1960. 10 P. Schiessl and M. Raupach, Laboratory Studies and Calculations on the Influence of Crack Width on Chloride-Induced Corrosion of Steel in Concrete, ACI Materials Journal, Vol. 94(1), JanuaryFebruary, 1997, p. 56. 11 Debate: Crack Width, Cover, and Corrosion, Concrete International, May, 1985, p. 20. 12 A.A. Sags, S.C. Kranc, F. Presuel-Moreno, D. Rey, A. Torres-Acosta, and L. Yao, Corrosion Forecasting for 75-Year Durability Design of Reinforced Concrete, Final Report No. BA502 submitted to Florida Department of Transportation by University of South Florida, December 31, 201. 13 Florida Method of Test for Determining Low-Levels of Chloride in Concrete and Raw Materials, FM 5-516, Florida Department of Transportation, September 1994. 14 N.S. Berke, M.P. Dellaire, M.C. Hicks, and R.J. Hoopes, Corrosion of Steel in Cracked Concrete, Corrosion, Vol. 49, 1993, p. 934. 15 F.J. Presuel-Moreno and D. Rourke, Towards Achieving the 100 Year Bridge Using Galvanized Rebar, paper presented at the International Corrosion Congress, Las Vegas, May, 2008. 16 P. Schiessel, Admissible Crack Width in Reinforced Concrete Structures, contribution no. II, InterAssociation Colloquium on the Behavior of Interservice Concrete Structures, Preliminary Report V.II, Liege, 1995, p. 3. 17 D. Darwin, J. Browning, M. OReilly, and L. Xing, Critical Chloride Threshold for Galvanized Reinforcing Bars, research report 07-2 submitted to the International Lead Zinc Research Organization, Inc. by University of Kansas, December, 2007. 18 Corrosion Resistant Alloys for Reinforced Concrete, Report No. FHWA-HRT-07-039, Federal Highway Administration, Washington, DC, July, 2007.

Page 10 of 10

S-ar putea să vă placă și