Sunteți pe pagina 1din 6

Submitted September 2004

Ultrasonic Testing of Concrete Structures


by John S. Popovics*

ABSTRACT
Although the application of conventional ultrasonic techniques are limited by the nature of concrete, modified ultrasonic techniques have been applied to detect internal cracking and other discontinuities, monitor changes such as deterioration due to aggressive environments, estimate inplace strength and monitor material strength and stiffness gain over time. This paper reviews the application of selected ultrasonic nondestructive testing (NDT) techniques to concrete structures. The theoretical bases and onsite application cases of the through thickness pulse velocity test, ultrasonic imaging and high frequency ultrasonic interface reflection measurements are described. Keywords: concrete, cracking, imaging, nondestructive testing, strength, ultrasonic pulse velocity.

INTRODUCTION
Structural concrete constitutes a significant portion of the total infrastructure in the United States. Much of this concrete infrastructure is approaching, or has passed, its original design life. Therefore, at present, it is in poor condition, as described by the Report Card of the American Infrastructure (American Society of Civil Engineers, 2003). In that report, it is estimated that a $1.3 trillion investment is needed to return the condition of the infrastructure to acceptable levels. Before appropriate rehabilitation can be prescribed, however, the condition of the structures must be assessed. Nondestructive testing (NDT) techniques that can detect, localize and characterize damage, deterioration and discontinuities in concrete are of great interest to infrastructure management agencies. This is important for concrete structures since internal discontinuities may remain hidden from view, yet significantly compromise the overall integrity of the structure. Early detection of discontinuities is the most effective way to reduce maintenance and rehabilitation costs, thereby improving public safety. The deterioration and damage of concrete structures can be described by a variety of physical and environmental damage modes. Common discontinuity modes include delamination/spalling, cracks and voids/honeycombs. In other cases, the geometry or size of a concrete element should be verified in a nondestructive fashion: for example, verifying the thickness of pavement slabs or the depth and shape of poured concrete shafts that act as the supporting foundations for tall buildings. Finally, the strength and stiffness development of early age concrete over time may be needed. Concrete structures present several specific challenges that must be overcome before NDT can be applied effectively. The structures usually are large in size. As a result, much time is needed to adequately test the entire structure and deep material penetration is often difficult to achieve. In addition, concrete naturally exhibits
* The University of Illinois, 2129D Newmark Laboratory, MC-250, 205 N. Mathews Ave., Urbana, IL 61801; (217) 244-0843; fax (217) 265-8040; e-mail <johnpop@uiuc.edu>.
50 Materials Evaluation/January 2005

large scale inhomogeneous material structure and a great deal of local material property variability. Structural concrete is composed of graded mineral aggregates, up to 30 mm (1.2 in.) in size, bound by an inorganic cement matrix. In addition, most structural concrete contains a grid of reinforcing steel bars or cables. For these reasons, many conventional NDT techniques that work well for steel and other homogeneous materials cannot be applied to concrete. Concrete is unable to transmit high frequencies, as the inhomogeneity of the concrete causes signals of smaller wavelengths or wavelengths equal to the nominal aggregate size to be highly scattered and attenuated. For example, high frequency pulse/echo A-scan or C-scan analysis cannot be applied directly to concrete because of the intensive backscatter caused by the aggregates. Nevertheless, some forms of ultrasonic tests have found application in concrete structures. (In this paper, ultrasonic tests are defined as dynamic measurements using wave frequencies of 20 kHz and greater, where the waves are generated by an electromechanical transducer. NDT that makes use of waves generated by an impact event, or sonic tests, will not be discussed here.) The development of through thickness ultrasonic wave velocity measurement, commonly referred to as the ultrasonic pulse velocity method, began in Canada and England as early as the 1940s. Since then, many nations have adopted standardized procedures to measure pulse velocity in concrete. In the 1960s, one sided pulse/echo systems were developed for concrete (Bradfield and Woodroffe, 1964) and over the next several decades, additional advancements in ultrasonic testing of concrete have been introduced, both for practical field application and for laboratory research purposes (Suaris and Fernando, 1987; Jacobs and Whitcomb, 1997; Sellick et al., 1998; Popovics et al., 2000; Purnell et al., 2004). Nevertheless, through thickness wave velocity tests have remained the most common ultrasonic tests and the only ones standardized for use in concrete. In this paper, three types of ultrasonic tests for concrete structures are described: the through thickness pulse velocity test, ultrasonic imaging and high frequency ultrasound interface reflection measurements.

ULTRASONIC PULSE VELOCITY


Background The ultrasonic pulse velocity test measures the travel time of ultrasonic P-wave (also called compressional wave or L-wave) pulses over a known path length using two ultrasonic transducers. To measure the ultrasonic pulse velocity, the sending transducer, driven by a generator, transmits a wave pulse into the concrete and the receiving transducer, separated from the sender by the distance L, receives the pulse through the concrete at another point. The wave pulse transmitted to the concrete undergoes scattering at various aggregate/cement binder boundaries. By the time the pulse reaches the receiving transducer, it gets transformed into a complicated

form, which contains multiple reflected P-waves and mode converted S-waves (shear waves). However, P-waves travel the fastest and thereby arrive first at the receiver. The pulse velocity instrument then measures this time of flight T for the first arriving P-wave pulse to travel through the concrete. Alternatively, T can be determined by direct testing of the received wave signal with respect to the known start time of the signal if the equipment is connected to an oscilloscope, or other display device, to observe the nature of the received wave signals. The P-wave pulse velocity V is then given by (1) V= L T (a)

The velocity of a wave depends on the elastic properties and density of the medium. For elastic, homogeneous solid media, the P-wave velocity is given by the following (American Concrete Institute, 1998) KE V= (2) where K = (1 )/[(1 + ) (1 2)] E = the modulus of elasticity = density = Poissons ratio. The value of K varies within a fairly narrow range. For example, as increases from 0.15 to 0.25 (67% increase), the associated K value increases from 1.06 to 1.20 (12% increase). Variations in E and have more significant effects on V than variations in . For concrete, V typically ranges from 3000 to 5000 m/s (9840 to 16 400 ft/s), depending on concrete strength, age, moisture content, type and amount of aggregate and proximity to steel reinforcing bars (Naik et al., 2004). Portable ultrasonic pulse velocity testing units have become available worldwide. Transducers with frequencies of 25 to 100 kHz are usually used for testing concrete, but transducer sets having different resonant frequencies are available for special applications. For concrete, the upper limit of usable frequency is about 500 kHz since the associated wavelength is in tens of millimeters, which is in the size range of the coarse aggregate particles. As a result, the path length that can be effectively traversed at this upper limit of frequency before the wave pulse becomes completely scattered is only several centimeters. Greater path lengths can be traversed using lower frequencies (and thus larger wavelengths); a frequency of 20 kHz can usually traverse up to 10 m (32.8 ft) of concrete (Jones, 1962). In order to transmit and receive the wave pulse, the transducers must be well coupled to the concrete; otherwise an air pocket between the transducer and test medium may introduce an error in the indicated transit time. Petroleum jelly has proven to be a good couplant for many cases. If the concrete surface is very rough, thick grease should be used as a couplant. In some cases, the rough surface may have to be ground smooth or a smooth surface may have to be established with the use of plaster of paris, suitable quick setting cement paste or quick setting epoxy mortar. There are three possible configurations in which the transducers may be arranged,as shown in Figure 1. These are: direct transmission (through thickness); semidirect transmission; and indirect or surface transmission. The direct transmission method is the most desirable and satisfactory arrangement because, with this arrangement, maximum energy of the pulse is transmitted and received. The semidirect transmission configuration can also be used quite satisfactorily. However, care should be exercised that the transducers are not too far apart, otherwise the transmitted pulse might attenuate and thus not be detected. The indirect or surface transmission configuration is least satisfactorily because the amplitude of the received signal is significantly lower. This method is also more prone to errors and a special procedure may be necessary for determining the pulse velocity (Qixian and Bungey, 1996; Benedetti, 1998).

(b)

(c) Figure 1 Testing configurations for the ultrasonic pulse velocity test (T = transmitting transducer, R = receiving transducer): (a) direct; (b) semidirect; (c) indirect (Naik et al., 2004).

FACTORS AFFECTING THE PULSE VELOCITY IN CONCRETE


Aggregate Size, Grading, Type and Content The pulse velocity is affected significantly by the type and amount of mineral aggregate contained in the concrete mixture (Kaplan, 1959; Anderson and Seals, 1981; Sturrup et al., 1984; Swamy and Al-Hamed, 1984; Popovics et al., 1990). In general, the P-wave velocity of the cement binder is lower than that of aggregate; for the same concrete mixture at the same compressive strength level, concrete with rounded gravel had the lowest pulse velocity, crushed limestone resulted in the highest pulse velocity and crushed granite had a velocity between these two. Research findings indicate that, at the same strength level, the concrete having the higher aggregate content gives a higher pulse velocity. The effects of varying the proportion of coarse aggregate in a concrete mixture on the pulse velocity versus compressive strength relationship are shown in Figure 2. The figure shows that for a given value of pulse velocity, the higher the aggregate to cement ratio, the lower the compressive strength. Moisture and Curing Condition of Concrete Concrete is a porous material and it will absorb externally supplied water into its internal pore structure. The pulse velocity for
Materials Evaluation/January 2005 51

Figure 2 Effect of cement to fine aggregate to coarse aggregate proportions on the relationship between pulse velocity and compressive strength (Naik et al., 2004). saturated concrete is higher than for dry concrete. A 4 to 5% increase in pulse velocity can be expected when dry concrete is saturated (Jones, 1962). Kaplan (1958) found that the pulse velocities for laboratory cured specimens were higher than for the site cured specimens. He also found that pulse velocity in columns cast from the same concrete were lower than in the site cured and laboratory cured specimens. Presence of Reinforcing Steel A significant factor that influences the pulse velocity of concrete is the presence of steel reinforcement. The P-wave velocity in steel is 1.4 to 1.7 times that in plain concrete. Therefore, pulse velocity readings in the vicinity of reinforcing steel are usually higher than those of plain concrete. Whenever possible, test readings should be taken such that the reinforcement is avoided in the wave path. If reinforcements cross the wave path, correction factors should be used. The importance of including bar diameters as a basic parameter in the correction factors has been demonstrated (Chung, 1978; Bungey, 1984). It should be emphasized, however, that in heavily reinforced sections it might not be possible to obtain accurate measurements of the concrete pulse velocity. Concrete Compressive Strength and Youngs Modulus The inplace compressive strength of concrete is an important design factor for structural engineers. Concrete gains strength over the first several months after casting as a result of a chemical reaction (the hydration reaction) between the inorganic cement binder and water. The strength of concrete is generally controlled by the relative amount of water to cement binder (the water to cement ratio by mass) in the concrete mixture and also the age of the concrete. As the water to cement ratio increases, the compressive and flexural strengths and the corresponding pulse velocity decrease, assuming no other changes in the composition of the concrete (Kaplan, 1958). The effect of concrete age on the pulse velocity is similar to the effect upon the strength development of concrete. The velocity increases very rapidly initially but soon flattens out (Jones, 1962). This trend is similar to the strength versus age curve for a particular type of concrete, but pulse velocity flattens sooner than strength. Once the pulse velocity curve flattens out, experimental errors make it impossible to estimate the strength with accuracy. Nevertheless, the pulse velocity method provides a means of estimating the strength of both in situ and precast concrete, although there is no physically based relation between the strength and velocity. The strength can be estimated from the pulse velocity by a preestablished graphical correlation between the two parameters, an example of which is shown in Figure 3. The relationship between strength and pulse velocity is not unique, as it is affected by other factors such as moisture content and aggregate size, type and content, as described above. Thus, no attempts should be made to
52 Materials Evaluation/January 2005

Figure 3 Example of a strength versus velocity relationship for estimation of concrete strength (Naik et al., 2004).

estimate compressive strength of concrete from pulse velocity values unless similar correlations have been previously established for the type of concrete under investigation. The American Concrete Institute (2003) provides recommended practices to develop the relationship between pulse velocity and compressive strength, which can be later used for estimating the in situ strength based upon the pulse velocity. The velocity of a P-wave traveling through an elastic material is uniquely defined by the elastic constants and density of the material by wave propagation theory (see Equation 2). Therefore, it is possible to compute the modulus of elasticity of a material if the ultrasonic pulse velocity is measured where the values of Poissons ratio and density are known or assumed. This approach has an advantage over another standardized technique, which make use of vibration frequencies, in that the testing is not restricted to specially shaped laboratory specimens. Nevertheless, the estimation of the dynamic modulus of elasticity in concrete from ultrasonic pulse velocity measurements is not normally recommended for two reasons: the error resulting from inaccurate estimation of Poissons ratio is not insignificant; and Equation 2 is appropriate for homogeneous materials only, leaving the validity for inhomogeneous composite materials, such as concrete, in doubt. Usually, the dynamic modulus of elasticity estimated from pulse velocity measurements is higher than that obtained from vibration measurements, even when the value of Poissons ratio is known (Philleo, 1955). Internal Discontinuities Cracks, voids and other internal discontinuities in concrete will cause variations in the measured pulse velocity. For example, the diffraction of a wave pulse around an internal air void will cause an increase in the time of propagation with respect to a path through the void center and the measured apparent velocity will decrease. This occurs because mechanical waves in solids are not readily transmitted across air or water interfaces. However, only large cracks and voids, generally larger than the transducer contact face, will cause measurable reduction in velocity (Jones, 1962). In this regard, the pulse velocity method is effective in establishing comparative data for qualitative testing of concrete and is suitable for the study of concretes heterogeneity. Here, heterogeneity is defined as interior cracking, deterioration, honeycombing and variations in mixture proportions. For obtaining these qualitative data, a system of measuring points, for example, a grid pattern, should be established. Depending upon the quantity of the concrete to be tested, the size of the structure, the variability expected and the accuracy required, a grid of 300 mm (11.8 in.) spacing, or greater, should be established. Generally about 1 m (39.4 in.) of spacing is adequate (American Concrete Institute, 1998). Other applications of this qualitative comparison of in situ or test specimen concrete are: to check the variation of concrete density in order to test the effectiveness of

consolidation; locating areas of honeycombed concrete; and localizing internal cracks and voids. Several researchers have applied ultrasonic pulse velocity to measure the depth of surface breaking cracks in concrete using the indirect testing configuration (Knab et al., 1983; Rebic, 1983). If a pulse traveling through the concrete comes upon an air filled crack or a void whose projected area perpendicular to the path length is larger than the area of the transmitting transducer, the pulse will diffract around the discontinuity, as illustrated in Figure 4. Thus, the pulse travel time will be greater than that through similar concrete without any discontinuity. It should be pointed out that the application of this technique in locating discontinuities has serious limitations. For example, if cracks and other discontinuities are small, if they are filled with water or other debris (thus allowing the wave to propagate through the discontinuity) or if the crack tip is not well defined, the pulse velocity will not significantly decrease, implying that no discontinuity exists (Naik et al., 2004).

or three dimensional property (for example, velocity) map of the structure within a specific region, indicated by the dotted line box in Figure 5. Tomographic imaging software must be used to reconstruct the data collected along ray paths at varying angles; the reconstruction process can be computationally intensive. Several different reconstruction algorithms may be applied, but sufficient data are needed in order to assure convergence to the correct solution (Gomm and Mauseth, 1998). Greater numbers of measurements and intersecting ray paths result in more accurate tomograms (Martin et al., 2001). A tomographic reconstruction of velocity data collected from a concrete beam containing two partially filled ducts is shown in Figure 6. Voids and poor concrete regions are indicated by low apparent velocity regions.

Figure 4 Scheme for measurement of surface breaking crack depth h (Naik et al., 2004).

ULTRASONIC IMAGING
Individual wave data characterize the material along a given wave path. Since the presence of discontinuities and variation in the internal material properties affect the travel time and amplitude of ultrasonic waves along each path, multiple exterior ultrasonic data projections may be assembled to build up an interior three dimensional image. Tomographic approaches and the synthetic aperture focusing technique enable the assembly of large amounts of data to form such images and have been applied to concrete structures (Rhazi et al., 1997). Both the tomographic and the synthetic aperture focusing technique require collection of a large amount of data to reconstruct adequate images, which is computationally and labor intensive. Tomography The tomographic technique combines large amounts of physical data, most often ultrasonic velocity values, taken from many different intersecting ray paths through the material, as illustrated in Figure 5. The projections are then used to reconstruct a cross sectional Figure 6 Velocity tomograph reconstructed with data from a reinforced concrete slab. Regions containing ducts, voids and poor concrete are identified as low velocity regions (Martin et al., 2001). Synthetic Aperture Focusing Technique The synthetic aperture focusing technique numerically superimposes many ultrasonic pulse/echo time signals, measured at several positions, to create a high resolution image. The signals are assembled and integrated, or focused, with respect to the time of flight surface in volume and time space for each voxel of material. Thus, structural noise is suppressed by spatial superposition (Schickert et al., 2003). Limitations to conventional ultrasonic A-scan testing, owing to the highly inhomogeneous nature of concrete, are thereby overcome. The synthetic aperture focusing technique result is a three dimensional map of the backscatter intensity from inside the structure, which may then be interpreted in area (C-scan) or depth (B-scan) slices (Krause et al., 2001). Indications of significant backscatter indicate locations of interior air filled voids and cracks. Reconstruction algorithms for one, two and three dimensional synthetic aperture focusing technique from contact pulse/echo measurements have been used to identify and locate indications of significant backscatter from backwall echoes, tendon ducts, voiding and steel reinforcement inside concrete (Schickert et al., 2003). Transducer arrays are usually used in the measurement. Figure 7 shows a three dimensional synthetic aperture focusing technique B-scan reconstruction, where upper and lower reinforcement layers perpendicular to the scanning direction and the backwall are imaged. At X = 860 mm (33.9 in.) and Z = 160 mm (6.3 in.), reflection from honeycombing is revealed, combined with shading of the backwall. A second honeycombing discontinuity is detected by shading of the backwall without direct indication (Y = 120 mm [4.7 in.]). Synthetic aperture focusing technique requires good coupling between transducers and test surface, which is often difficult to
Materials Evaluation/January 2005 53

Figure 5 Data collection scheme for tomographic reconstruction.

Figure 7 Three dimensional synthetic aperture focusing technique B-scan of a concrete slab containing steel reinforcement and two artificial discontinuities (Schickert et al., 2003).

apply to practical testing of concrete due to the inherently rough surface of concrete. In addition, only larger discontinuities (greater than 200 mm [7.9 in.] in size) can be detected reliably (Popovics, 2003).

INTERFACIAL WAVE REFLECTION


Monitoring the setting and hardening process (stiffness change over time owing to the hydration of the cement binder) in concrete is often of importance with respect to economy and structural safety. For example, the efficiency of precast concrete element production can be raised or the time for concrete form removal can be optimized. Investigators have reported nondestructive monitoring of the hydration process of cement, including several that use ultrasound for this purpose. Initial efforts applied ultrasonic through transmission measurements to characterize the development of the mechanical properties of concrete (Boumiz et al., 1996; Arnaud, 2003). More recently, it has been shown that an ultrasonic interfacial wave reflection technique has significant advantages compared to the through transmission technique. The test can be applied to structures allowing limited access and data collection can be started immediately after mixing and continue indefinitely. Several researchers have shown that these wave reflection data are directly connected to the mechanical properties of the stiffening concrete, including its strength development (ztrk et al., 1999; Valic, 2000; Voigt et al., 2003). In the technique, an ultrasonic wave pulse (usually 1 to 5 MHz) is launched in a steel plate in contact with the concrete and reflections from the steel/concrete interface are monitored. The wave reflection method monitors the development of the reflection coefficient at the interface between steel and concrete over time, thus the stiffness change (setting) of the concrete is inferred. When a wave encounters the steel/concrete interface, part of the wave energy is transmitted into the concrete and part is reflected back to the transducer. Some of this wave energy is then again reflected from the transducer/steel interface into the steel and is again partially reflected when it again hits the steel/concrete boundary. The process is illustrated in Figure 8 where ST is the transducer signal transmitted into the steel, R1 and R2 are the first and second reflections captured by the transducer and T1 and T2 are the first and second transmission into the concrete. When a P-wave or S-wave is reflected at a boundary between two different materials, the reflection coefficient r can be calculated as (3) r= 2 v 2 1 v1 2 v 2 + 1 v1

Figure 8 Schematic representation of the reflection and transmission process at the steel/concrete interface (Voigt et al., 2003). The development of concretes shear modulus in particular is related to how the microstructure of hydrating cement binder evolves as a result of curing, indicating the development of compressive strength. Thus, S-waves are often used since the reflection coefficient so determined is governed by the development of the concrete shear modulus. An S-wave traveling through metal that is incident upon a steel/fluid interface is entirely reflected. Thus, at early ages most of the S-wave energy is reflected and the amplitude of the received wave is large. As the concrete stiffens, more of the wave energy is transmitted through the concrete and less is reflected at the interface. The magnitude wave reflection can be quantified using a wave reflection factor, which defines the ratio of the amount of incident wave energy that is reflected from an interface between two materials (ztrk et al., 1999) or the same ratio expressed in decibels, called reflection loss (Voigt, 2003). Experimental tests results show that certain features in the reflection coefficient development curves of concrete correlate well to pin penetration tests and concrete temperature measurements. Plots of typical S-wave reflection loss (expressed in decibels) as a function of time are shown in Figure 9. The reflection is measured continuously after casting up to 72 h. Significant changes in the early response of the reflection factor coincide with distinctive stages of the hydration process. In the initial stage, the temperature inside the concrete decreases followed by a period of thermal inactivity, the dormant period. The reflection loss during this stage remains constant near zero. During this dormant period, the concrete mixture has considerable plasticity that is maintained for several

where 1 = the density of material 1 2 = the density of material 2 v1 = the wave velocity in material 1 v2 = the wave velocity in material 2.
54 Materials Evaluation/January 2005

hours, allowing mixing casting and finishing operations to be carried out. At the beginning of the acceleration period, the temperature inside the specimen increases due to the exothermal reaction of the cement coinciding with initial stiffening and setting. After 5 h, the concrete begins to stiffen noticeably, corresponding with the end of the induction period. At this time there is a noticeable increase in the reflection loss response away from zero. This point has been shown to correlate well with the time of initial setting. Afterwards, there is a steady, almost linear increase in the reflection loss, indicating that the observed trends are owing to the change in the mechanical properties of concrete. Inferences about concrete strength development are made during this stage.

Figure 9 S-wave reflection loss development for three different concrete batches (Voigt et al., 2003).

CONCLUSION
Despite the limitations imposed by the material character, ultrasonic tests have been applied successfully to concrete structures. The ultrasonic pulse velocity method is an effective and practical means of investigating the uniformity of concrete and, in conjunction with imaging procedures, provides an effective method to locate interior discontinuities within concrete. The method can also be applied to estimate the in place concrete strength; however, a large number of variables affect the relations between the strength parameters of concrete and its pulse velocity. The use of ultrasonic pulse velocity to estimate the compressive or flexural strengths of concrete is not recommended unless previous correlation testing has been performed. The ultrasonic interfacial wave reflection method is effective in monitoring the stiffening and hardening process of early age concrete.

REFERENCES
American Concrete Institute, ACI Committee 228, ACI 228.2R-98: Nondestructive Test Methods for Evaluation of Concrete in Structures, Farmington Hills, Michigan, American Concrete Institute, 1998. American Concrete Institute, Committee 228, ACI 228.1R-03: In-Place Methods to Estimate Concrete Strength, Farmington Hills, Michigan, American Concrete Institute, 2003. American Society of Civil Engineers, Report Card for Americas Infrastructure, <www.asce.org/reportcard/index.cfm>, Reston, Virginia, American Society of Civil Engineers, 2003. Anderson, D.A. and R.K. Seals, Pulse Velocity as a Predictor of 28 and 90 Day Strength, ACI Journal, Vol. 78, 1981, p. 116. Arnaud, L., Rheological Characterization of Heterogeneous Materials with Evolving Properties, ASCE Journal of Materials in Civil Engineering, Vol. 15, 2003, pp. 255-265. Benedetti, A., On the Ultrasonic Pulse Propagation into Fire Damaged Concrete, ACI Structural Journal, Vol. 95, 1998, p. 259. Boumiz, A., C. Vernet and F. Cohen Tenoudji, Mechanical Properties of Cement Pastes and Mortars at Early Ages, Journal of Advanced Cement-based Materials, Vol. 3, 1996, pp. 94-106. Bradfield, G. and E.P.H. Woodroffe, Determining the Thickness of Concrete Pavements by Mechanical Waves: Diverging Beam Method, Magazine of Concrete Research, Vol. 16, 1964, pp. 4563.

Bungey, J.H., The Influence of Reinforcement on Ultrasonic Pulse Velocity Testing, ACI SP 82, Farmington Hills, Michigan, American Concrete Institute, 1984, p. 229. Chung, H.W., Effect of Embedded Steel Bar upon Ultrasonic Testing of Concrete, Magazine of Concrete Research, Vol. 30, 1978, p. 19. Gomm, T.J. and J.A. Mauseth, State of the Technology: Ultrasonic Tomography, Materials Evaluation, Vol. 57, 1998, pp. 747-752. Jacobs, L.J. and R.W. Whitcomb, Laser Generation and Detection of Ultrasound in Concrete, Journal of Nondestructive Evaluation, Vol. 16, 1997, pp. 5765. Jones, R., Non-destructive Testing of Concrete, London, Cambridge University Press, 1962. Kaplan, M.F., Compressive Strength and Ultrasonic Pulse Velocity Relationships for Concrete in Columns, ACI Journal, Vol. 29, 1958, p. 675. Kaplan, M.F., The Effects of Age and Water to Cement Ratio upon the Relation between Ultrasonic Pulse Velocity and Compressive Strength of Concrete, Magazine of Concrete Research, Vol. 11, 1959, p. 85. Knab, L.J., G.V. Blessing and J.R. Clifton, Laboratory Evaluation of Ultrasonics for Crack Detection in Concrete, ACI Journal, Vol. 80, 1983, p. 17. Krause, M., F. Mielentz, B. Milman, W. Muller, V. Schmitz and H. Wiggenhauser, Ultrasonic Imaging of Concrete Members Using an Array System, NDT&E International, Vol. 34, 2001, pp. 403-408. Martin, J., K.J. Broughton, A. Giannopolous, M.S.A. Hardy and M.C. Forde, Ultrasonic Tomography of Grouted Duct Post-tensioned Reinforced Concrete Bridge Beams, NDT&E International, Vol. 34, 2001, pp. 107-113. Naik, T.R., V.M. Malhotra and J.S. Popovics, CRC Handbook for Nondestructive Testing of Concrete, second edition, N.J. Carino and V.M. Malhotra, eds., Boca Raton, Florida, CRC Press, 2004. ztrk, T., J. Rappaport, J.S. Popovics and S.P. Shah, Monitoring the Setting and Hardening of Cement-based Materials with Ultrasound, RILEM Concrete Science and Engineering, Vol. 1, 1999, pp. 83-91. Philleo, R.E., Comparison of Results of Three Methods for Determining Youngs Modulus of Elasticity of Concrete, ACI Journal, Vol. 26, 1955, p. 461. Popovics, J.S., NDE Techniques for Concrete and Masonry Structures, Progress in Structural Engineering and Materials, Vol. 5, 2003, pp. 49-59. Popovics, S., J.L. Rose and J.S. Popovics, The Behavior of Ultrasonic Pulses in Concrete, Cement and Concrete Research, Vol. 20, 1990, p. 259. Popovics, S., N.M. Bilgutay, M. Karaoguz and T. Akgul, High-frequency Ultrasound Technique for Testing Concrete, ACI Materials Journal, Vol. 97, 2000, pp. 58-65. Purnell, P., T.H. Gan, D.A. Hutchins and J. Berriman, Noncontact Ultrasonic Diagnostics in Concrete: A Preliminary Investigation, Cement and Concrete Research, Vol. 34, 2004, pp. 1185-1188. Qixian, L. and J.H. Bungey, Using Compression Wave Ultrasonic Transducers to Measure the Velocity of Surface Waves and Hence Determine Dynamic Modulus of Elasticity for Concrete, Construction and Building Materials, Vol. 10, 1996, p. 237. Rebic, M.P., The Distribution of Critical and Rupture Loads and Determination of the Factor of Crackability, ACI Journal, Vol. 80, 1983, p. 17. Rhazi, J., Y. Kharrat, G. Ballivy and M. Rivest, Application of Acoustical Imaging to the Evaluation of Concrete in Operating Structures, ACI SP 168, Farmington Hills, Michigan, American Concrete Institute, 1997, p. 221. Schickert, M., M. Krause and W. Muller, Ultrasonic Imaging of Concrete Elements Using Reconstruction by Synthetic Aperture Focusing Technique, ASCE Journal of Materials in Civil Engineering, Vol. 15, 2003, pp. 235-246. Selleck, S.F., E.N. Landis, M.L. Peterson, S.P. Shah and J.D. Achenbach, Ultrasonic Investigation of Concrete with Distributed Damage, ACI Materials Journal, Vol. 95, 1998, p. 27. Sturrup, V.R., R.J. Vecchio and H. Caratin, Pulse Velocity as a Measure of Concrete Compressive Strength, ACI SP 82, Farmington Hills, Michigan, American Concrete Institute, 1984, p. 201. Suaris, W. and V. Fernando, Ultrasonic Pulse Attenuation as a Measure of Damage Growth during Cyclic Loading of Concrete, ACI Materials Journal, Vol. 84, 1987, pp. 185-193. Swamy, N.R. and A.H. Al-Hamed, The Use of Pulse Velocity Measurements to Estimate Strength of Air-dried Cubes and Hence In Situ Strength of Concrete, ACI SP 82, Farmington Hills, Michigan, American Concrete Institute, 1984, p. 247. Valic, M.I., Hydration of Cementitious Materials by Pulse Echo USWR Method, Apparatus and Application Examples, Cement and Concrete Research, Vol. 30, 2000, pp. 1633-1640. Voigt, T., Y. Akkaya and S.P. Shah, Determination of Early-age Mortar and Concrete Strength by Ultrasonic Wave Reflections, ASCE Journal of Materials in Civil Engineering, Vol. 15, 2003, pp. 247-254.

Materials Evaluation/January 2005 55

S-ar putea să vă placă și