You are on page 1of 21
Flexible Pavement Analysis 7.1 Introduction Flexible pavements are modeled as layered elastic systems with infi- nite lateral dimensions. These layers rest on the subgrade, which is often modeled as an elastic layer of infinite depth. Elasticity implies that all the pavement layers and the subgrade can be described by their elastic Young’s modulus E and their Poisson's ratio 4. Further- more, the layers are assumed to be homogeneous and isotropic. Tire loads are modeled as either point loads or circular loads of uniform pressure using Equation 2.1. Under these conditions, the stress state is axisymmetric; that is, it exhibits rotational symmetry around the center axis of the load and, as a result, it is easier to describe using a radial coordinate system. Pavement responses, (i.c., stress, strains, and deflections) are calculated using relationships from the theory of elasticity. The responses from multiple loads are calculated by superimposing the stresses from the individual tires, according to D’Alembert’s superposition principle. Analyzing these responses is essential for the mechanistic design of asphalt concrete pavements, as described in Chapter 11. The following discussion describes solutions for single-layer, two-layer and multilayer flexible pavement systems. The granular layers are treated as linear elastic and nonlinear elastic, (i.¢., having 183 184 7 Flexible Pavement Analysis moduli that are stress-independent and stressdependent, respec- tively), while the asphalt concrete is treated as either linear elastic or linear viscoelastic. 7.2 Single-Layer Elastic Solutions 7.2.1 Point Load The simplest loading condition is that of a single-point load, P, applied on a semi-infinite elastic space illustrated in Figure 7.1, which shows the nonzero stresses at a location defined by a depth zand a radial offset r. Obviously, due to axial symmetry, the radial location defined by the angle @ is not relevant. The stresses are defined as: o, = vertical normal stress o,= radial normal stress a9 = tangential normal stress Tz, = horizontal shear stress in the radial direction vy Figure 7.1 Axisymmetric Stresses State in an Elastic Half Space 7.2 SingleLayer Elastic Solutions The corresponding strains are: €, = vertical normal strain radial normal strain ea Yer angential normal strain horizontal shear strain in the radial direction. In defining these strains, it should be noted that the displacement field is two-dimensional; that is, a point in this semielastic space can move only vertically or horizontally, denoted by wand u, respectively, as shown in Figure 7.1. Hence: (7.1a) (7.1b) £9 =" (7.10) yon St (7.1d) where Equation 7.1c suggests that the tangential normal strain is, in essence, the change in the perimeter of the circle with radius + divided by the original perimeter, that is, [2x (r+ u) — 2ar]/(2r7). The closed-form solution to this problem was originally developed by Boussinesq, circa 1880", and adapted by Taylor!? in the following form: Pp Qn (P+ (7.2a) ys (7.2b) P 39°. 1-2u o, =—-2- | i l(2yey” Peetweee P Og = — (1— 2p) oor la 1 SS (7.20) Peet wl? + 5] (7.2) 185 186 7 Flexible Pavement Analysis Note that for normal stresses, the sign notation is minus for com- pression and positive for tension. Note also that directly under the point of load application, (i.c., r = 0, z = 0), the stresses are undefined. The strain components can be calculated from the stress components through generalized Hoek’s law. 1 aah (0: — Loy + 98)) (7.3a) 1 E (0, — w(o. + 08) (7.3b) 1 £9 = (04 — KL (0, + 9:)) (7.3c) Qty (+H) te G (7.3d) where G is the shear modulus of the elastic medium. These stress-strain relationships can be written in matrix form as: oO d-w) mu BX 0 ey o, E hw (-H) wb 0 er oof A+md—2)} » nh (=n) 0 £4 Tey 0 0 0 lle (74) The vertical and horizontal deflections, w and u, at any point, are computed by integrating the vertical and horizontal strains, respectively. The resulting expressions are: on faswe (+2)? 420-24) (2 +2) '") (75a) pA tH) A= 2H) ca sey a4 tae +2" QnrE 1-2 (7.5b) It should be noted that the surface (ie., z deflection is: P(l-w Pow) (7.6) Er 0), the vertical This expression defines the so-called solid or Boussinesq subgrade foundation model for rigid pavements, which is discussed further 7.2 Single-Layer Elastic Solutions 187 in Chapter 8. Using these closed-form stress, strain and deflec- tion expressions are straightforward, as explained in the following example. Compute the stresses and strains from a point load of 40 kN resting Example 7.1 ona semi-infinite elastic space. The location of interest is at a depth of 0.1 mand a radial offset of 0.2m. Given, E = 140 MPaand yu = 0.4. ANSWER: Substituting the specified values into Equation 7.2 gives the stresses: 3 = 20801 —34.16kPa 2m (0,22 + 0.12)" 40 30.2°0.1 1-204 oo | DEE | = 119.06 kPa ax [i 240.12) a a 0.1 1 40 oo = 5 (1 — 20.4) | °° Oe lei 0.22 +0.12 +. 0.10.22 + 0.1 | = —6.21 kPa Ter = #00201" se soura 2m (0,22 + 0.12)" Equation 7.3 gives the strains: = ra (34.16 — 0.4 (-119.06 — 6.21)) = 113.9 107% &= eas (—119.06 — 0.4 (—34.16 — 6.21)) = —735 10- o= aes (6.21 — 0.4 (-119.06 — 34.16) = 393 10° Ver = piace a OA) _ 1367.2 107° The response under a uniformly distributed stress p on a perfectly 7.2.2 Circular flexible circular area of radius a (e.g., an idealized tire imprint, as Load with Uniform defined in Chapter 2) is obtained by integrating the stress compo- _ Vertical Stress nents given by Equation 7.2. For points on the centerline of the load 188 Example 7.2 7 Flexible Pavement Analysis s expressions are!!: (7.7¢) and the vertical deflection under the centerline of the load is given by: _ 2-4 w= (7.8) Compute the stresses from a tire inflated to 600 kPa, carrying 30 kN resting on a semi-infinite elastic space. The location of interest is ata depth of 0.1 m and a radial offset of 0.0m. Also, compute the surface deflection, (i. 0.00) under the same tire. Given, E 140 MPa and p= 0.4. ANSWER: The radius of the tire imprint is computed from Equation 2.1. — 30 =0.126 m 600 7 ‘The stresses are computed by substituting the given data into Equation 7.7. B oz = 600} —1 + ——____, | = 455.9 kPa (0.126? + 0.12) 600 2(1 +0.4)0.1 0.1 0, = 09 = +2044 2020801 or 2 0.1267 + 0.17 — (0.1962 + 0.12)” = —89.9kPa 7.3 TwoLayer Elastic Solutions 189 The surface vertical deflection is computed from Equation 7.8: 20-048) 6 0 sop 5 w= 740000 600 0.126 = 0.907 10 7.3 Two! -Layer Elastic Solutions _ This system consists of a finite thickness layer placed of on top of another layer of infinite thickness. These two layers have different clastic properties, as shown in Figure 7.2. This is an idealized representation of a simple pavement consisting of a stiffer layer (e.g., asphalt concrete) resting on a weaker foundation (i.e., subgrade). Burmister! developed the solution for the surface deflection of this system under the centerline of a uniform vertical stress p distributed over a circular area of radius a, assuming a Poisson’s ratio of 0.5. In condensed form, this is expressed as: where F,, is a function that depends on the ratios a/h and E2/E, where h is the thickness of the finite layer. Burmister produced a chart for F, for selected ratios of 4/h and Eg/E} based on the theory of elasticity (Figure 7.3). Figure 7.2 Schematic of Two-Layer Elastic System. 190 7 Flexible Pavement Analysis Lop + 2 3 4 3 6 7 8 ° 0 oo > 08 A. [ee Az 2ae=aneeaee ” Sa pi ee DS eS pr eT oe x ae - SS —| TI p++ } 1 — fe 2a— > a Ee 1 ral ; I o1 FE cs setts sing Lnes °, wo on FF Motus =) Suge Lae? Figure 7.3 r 2 3 4 5 6 7 8 9 10 Ratio ofthe Radius of the Flexible Bearing Area tothe Thickness of the Reinforcing Layer-1_ allt Fy Factors for Computing Surface Deflection at the Centerline of a Circular Imprint Carrying Uniform Stress (Ref. 1 Reproduced by Permission) Example 7.3 It is worthwhile to note the form similarities between Equation 7.9 and 7.8 when substituting a 4 value of Compute the surface deflection from a circular tire imprint with 0.1m radius carrying a pressure of 700 kPa resting on a 0.2m thick asphalt concrete layer placed on top of a subgrade of infinite depth. ‘The layer moduli are 1400 MPa and 140 MPa, respectively, and x is 0.5 for both layers. ANSWER For a/h = 0.5 and Ey/E, = 0.1, Figure 7.3 gives which substituted into Equation 7.9 gives: Fy value of 0.32, _ 15 700 0.1 - = 3 5 = Tq0000 032 = 0.24 107m or 0.24 mm, Fox? developed expressions for the stress components in the two-layer system, which were subsequently implemented through nomographs (e.g., reference 2). Advent of modern computers, how- ever, enabled solving this problem for multiple layers, as described 7.3 TwoLayer Elastic Solutions 191 next, rendering such nomographs obsolete. Itis, nevertheless, worth- while mentioning an approximate method for solving the elastic layered system problem attributed to Odemark*®. It consists of uanslating multiple layers of different moduli into an equivalent single layer, hence known as the method of equivalent thicknesses. For a system of two layers, such as the one shown in Figure 7.2, the top layer with thickness / can be translated into an equivalent thickness /,, with a modulus £9. For 4) = 2, the equivalent layer thickness of the top layer is given by: \ YS hy =00(2) h (7.10) B where 0.9 is an approximation factor. This allows utilizing the single-layer solutions in computing pavement responses in the lower layer. Compute the stresses at the bottom of a flexible pavement surface Example 7.4 layer 0.3 m thick resting on a semi-infinite subgrade layer. The load consists of a circular tire with a 0.1m radius carrying a uniform pressure of 700kPa. The stresses are to be computed under the centerline of the load. The layer moduli are 1400 MPa and 140 MPa, respectively, and 11 is 0.5 for both layers. ANSWER Equation 7.10 gives the equivalent thickness of the top layer in terms of the modulus of the bottom layer as: 1 1400) * h. = 0.9{ ——} 0.3 = 0.58! he o0( AR) 0 0.582 m. which allows using Equation 7.7 to compute the stresses in the subgrade, At the bottom of the top layer, they are: 5893 140582 _ | = 99.9 kPa (0.12 +0.5822)°" —0.218 kPa 2 5)0.589 5893 [-« 4205) 4 2(1 + 0.5)0.582 0.582 | VOIP + 0.582 — (0.12 + 0.5822) 192 7.4 Multilayer Linear Elastic Solutions 7 Flexible Pavement Analysis This approach can be extended to translate the thickness of multiple layers # to an equivalent thickness with a modulus equal to the modulus of the bottom layer Ey. he -1Z {a where f is the value of an approximation factor that depends on the ratio of the layer moduli and the relative magnitude of the layer thicknesses, with respect to the radius of the load. (7.11) A multilayer elastic system consists of multiple finite-thickness layers resting on a subgrade of infinite thickness (Figure 7.4). It is an ide- alized representation of multiple pavement layers, such as asphalt concrete friction and leveling layers, base layers and sub-base layers, each having different elastic properties. Elastic response solutions to this system were developed by extending the Burmister analytical approach for the two layer system (1) to multiple layers. A variety of software implements such solutions, such as ELSYM5’, DAMA®, KENLAYER®, and EVERSTRESS®. These programs have, to a great Figure 7.4 ‘Schematic of Multilayer Elastic System 7.4 Multilayer Linear Elastic Solutions extent, similar features and input requirements and can handle up to five layers, including an infinite depth subgrade. They accept multiple circular tires and compute stresses at any location in the layered system. The calculations are made in the radial coordinate system, and then translated to a Cartesian coordinate system, with its origin in the middle of a tire imprint (Figure 7.4). In the following examples, the computer program EVERSTRESS 5.0° is used. Sug- gestions on available layered analysis software and their sources are given on the Web site for this book, at www.wiley.com/go/pavement. Compute the stresses, strains, and deflections in a layered system under a set of four tires as shown in Figure 7.5. The locations of interest are the bottom of the asphalt concrete layer and the top of the subgrade, at coordinates (x = 0, y = 0). Assume full friction between the layers. ANSWER ‘The output of EVERSTRESS is shown in Table 7.1, which lists stresses, strains, and deflections at the locations requested. They 12m, 3000s Subgrade Base = M005 Figure 7.5 Layered Elastic System for Example 7.7 193 Example 7.5 194 Table 7.1 fas ‘of Problem 7.5: Responses under (x = 0, y = 0) Computed with EVERSTRESS 7 Flexible Pavement Analysis were obtained by setting the coefficient of friction between the layers to 1.0. As discussed later in Chapter 11, traditional flexible pavement mechanistic design is based on controlling two of these response parameters, namely the tensile strain at the bottom of the asphalt concrete layer and the compressive strain at the top of the subgrade. They are associated with bottom-up fatigue cracking and subgrade plastic deformation (i.e., rutting), respectively. As shown. in Table 7.1, the magnitude of the asphalt concrete bottom tensile strains is 101.4 10-® and 141.8 10-® in the transverse and longitudinal directions, respectively, while the subgrade top compressive strain is 166.1 10-®. It should be noted that the maximum strain under this set of tires may not be under the (x = 0, y = 0) location. 7.5 Multilayer Nonlinear Elastic Solutions As discussed in Chapter 3, granular materials exhibit a stress- dependent behavior; that is, their resilient (ie., elastic) modulus is a function of the stress state at any given location. Coarse-grained layers (e.g., base, subbase) exhibit an exponential dependence to bulk stress, while fine-grained/cohesive layers (e.g., subgrade) exhibit an exponential dependence to deviatoric stress (Equations 3.4 and 3.8, respectively). Layer elastic analysis can handle this nonlinearity using a piecewise linear iterative algorithm. Seed moduli are initially assigned; the layered analysis is conducted to compute initial stresses; and in subsequent iterations, the moduli are updated on the basis of the calculated stresses, as schematically shown in Figure 7.6. The process 7.5 Multilayer Nonlinear Elastic Solutions 450,000 400,000 350,000 4 300,000 4 250,000 EkPa Sood 100,000 50,000 0 ° 200 400, 600 800 1000 theta kPa Figure 7.6 ‘Schematic Representation of the Iterative Procedure Used to Handle Granular Layer Nonlinearity is repeated until the computed moduli in two successive iterations are within a prescribed tolerance. This process can be refined by subdividing each granular layer into sublayers and considering the stresses in the middle of each sublayer. In doing so, the weight of the layers should also be considered, EVERSTRESS* handles nonlinear granular material moduli. Repeat the solution of the previous example by considering the stress dependence of the base/subbase and subgrade layers given by: E=4000°% (7.12) E = 120(o; —03)°? (7.18) respectively, where E is the elastic modulus in MPa and @ is the bulk stress in atmospheres (ie., 01 + 62 + 93), and 04, 02, and a are the principal stresses in atmospheres. The unit weight of the pavement layers are 22.8, 20.5, 19.7, and 18.9 kN/m? (i.e., these are used for computing the stress from the overburden to be added to the load-induced stresses to obtain the bulk and deviator stresses). 195 Example 7.6 196 Table 7.2 ale of Example 7.6: Responses under (x = 0, y = 0) Computed with EVERSTRESS ef. 7 Flexible Pavement Analysis ANSWER: The answer is given in Table 7.2, which shows the magnitude of the asphalt concrete bottom tensile strains is 138.4 10-® and 194.1 10°® in the transverse and longitudinal directions, respectively, while the subgrade bottom compressive strain is 238.1 10~°, The solution was achieved in seven iterations, satisfying a tolerance in E lower than 1%. After convergence, the moduli of the base, subbase and subgrade layers were 374.3 MPa, 238.61 MPa and 88.31 MPa, respectively, 7.6 Viscoelastic Solutions The previous discussion assumed elastic material behavior. As described in Chapter 5, however, asphalt concretes exhibit vis- coelastie behavior, hence their response is time-dependent. Their response to a time-dependent (e.g., moving) load is computed through Bolumann’s superposition principle!®, assuming linear vis- coelastic behavior. In the time domain, this is expressed by the following convolution integral: : ew= [o-8) 2) ag ray where, ¢ (1) is the strain at time ¢, D(t — &’) is the creep compliance or retardation modulus of the asphalt concrete layer after a lapsed time of (t—£’), and o(&’) is the stress history as a function of time 7.6 Viscoelastic Solutions &' ranging from 0 to ¢, Plainly stated, Equation 7.14 means that the strain at time ¢, under an arbitrary stress history 0 (&’), is the linear sum of all the strain increments experienced from the beginning of the stress imposition to time ¢. The creep compliance is the inverse of modulus and has units of 1/stress (i.e. 1/kPa or 1/Ibs/in.2). Given a mechanistic model for the creep compliance, the integral in Equation 7.14 can be computed numerically, as demonstrated in the following example. It should be noted that this convolution integral can be written similarly in terms of stress rather than strain, expressed as: owe fee-9) Dae (7.15) where, (1) is the stress at time t, E(t — &') is the stiffness or relax- ation modulus of the asphalt concrete layer, and ¢ (&’) is the applied strain history. Mechanistic models can be fitted to describe the creep compliance or the relaxation modulus based on data obtained from creep testing or relaxation testing, respectively. Hence, the vis- coelastic behavior of asphalt binders and asphalt concretes can be effectively modeled through state-of-the art laboratory testing. Implementing Equation 7.14 in a semi-infinite elastic space is straightforward, because of the availability of closed-form solutions for the stress (Equation 7.2). An example of computing the response of a semi-infinite viscoelastic layer to a moving load is given in the following example. Consider a 40 KN point load traveling at a speed of 22.22m/sec (80 km/h) on a semi-infinite viscoelastic layer with a creep compli- ance given by: 1 Di) = (7.16) y+ Ret + Be ® with Zo = 5,000,000 kPa, E1 = 2,000,000 kPa, By = 4,000,000 kPa, 7) = 0.05 sec, and Ty = 0.005 sec. Compute and plot the vertical strain \s a function of time at a depth of 0.1m, assuming a constant Poisson’s ratio of 0.40. 197 7.6.1 Single Semi-Infinite Layers Example 7.7 198 7.6.2 Multilayer Systems 7 Flexible Pavement Analysis ANSWER A load influences layer response at a particular location before and after it reaches this location. The total distance of load influence depends on the magnitude of the load and the stiffness of the layer. For this example, assume that the total length of load influence is 1.0m (i.e., the response is substantial for locations of the load 0.5m before and 0.5 m after a point of interest). At 22.22 m/sec, the load takes 0.045 sec to traverse this 1.0m length of load influence (i.c., 1.0/22.22). Obviously, the resulting loading frequency is 22.22 Hz. According to Equation 7.3a, the vertical strain ¢, can be computed from the stress history of the three stress components @, 0, and 6. These can be computed from Equations 7.2 for the given depth z of 0.10m and any desired offset r ranging from —0.5 to 0.5 m (note that only stresses for the positive or the negative offsets need to be computed due to axial symmetry). Hence, according to Equation 7.14, and given a constant Poisson’s ratio, the vertical strain can be computed using: + (E') ~ 4 (or (6") +00 (&'))) ab" (7.17) This equation was solved numerically using a spreadsheet and time/distance increments of 0.00045 sec/0.01 m, as shown in Table 7.3. Note that the value of the creep compliance is com- puted from Equation 7.16 for values of time, increasing from 0 to 0.045 sec, while the stress history is calculated by positioning the 40 KN point load at successive locations 0.01 m apart and computing the change in the stress components. The resulting vertical strain versus time is plotted in Figure 7.7 and has a maximum compressive value of about 330 10-®. Modeling the response of layered systems under moving loads is more complex because there are no closed-form solutions for the stress components, hence layered analysis software has to be used to compute them (e. EVERSTRESS). It is not practical, however, doing so for every distance/time increment (it would involve 100 sets of analysis for the example just presented, with different surface layer moduli for each set). Alternatively, pavement response can be 199 LL ajdwex3 10} uojesZazuy jeouewny aiqeL 200 Example 7.8 7 Flexible Pavement Analysis z 2 | Ls06 0S 4 03-02 Ol 0 O01 02 03 04 05 Radial Offset (m) Figure 7.7 Vertical Strain 1 versus Time for Example 7.7 computed at selected locations and values in between approximated through interpolation, thus making it possible to use the same approach followed in the previous example. Consider a pair of identical tires moving at a speed of 16.667 m/sec (ie. 60km/h) on a pavement structure, as shown in Figure 7.8. Assume that these loads influence the response of the pavement within a radius of 1.00 m around their instantaneous location, (i.e., load influence length of 2.0m). Compute and plot the transverse and longitudinal strains (ie., ¢..and éy) at the bottom of the asphalt concrete layer as a function of time. 67 m/sec, the load takes 2.0/16.667 = 0.12sec to traverse its influence length (i.e., a loading frequency of 8.33 Hz). The stresses were computed at the bottom of the asphalt concrete layer (i.c., 0.152 m) in Cartesian coordinates using EVERSTRESS at y intervals of 0.10 m for the pavement layer layout shown in Figure 7.8. Stress values between these points were obtained through linear interpola- tion to yield stress estimates at increments of 0.01 m. Table 7.4 shows excerpts of the spreadsheet used to compute the stress history and the creep compliance values necessary for integrating numerically 7.6 Viscoelastic Solutions 201 1 Do= - : 250000 + 700000 ©" +2000000 « 689 kPa 0.096 m or o1stm) ac Ine 040, 04s? Base 413 MPa 0.40 | Suibgrade 103 as ve Figure 7.8 Layer Viscoelastic System for Example 7.8 Table 7.4 Numerical Integration for Example 7.8 Equation 7.17. The resulting transverse and longitudinal strains are plotted in Figure 7.9. An approximate method was proposed for computing pavement response under moving loads. It involves direct superposition of 202 7 Flexible Pavement Analysis ‘Strain (microns) Radial Offset (m) Figure 7.9 Transverse and Longitudinal Strains (ey, and ey) versus Time for Example 7.8 strains weighed by a loading function', expressed as: i alg" Rw = [re §') a Das (7.18) where R(t) is the pavement response parameter at time ¢ (e.g., the transverse strain éy, at the bottom of the asphalt concrete layer), I(Z) is the influence function for that pavement response parameter, and L(t) is the loading function. For a moving load of constant magnitude A, a common form for the loading function L(1) is one pulse-shaped o(™ mt D D L(t) = Asin® ( = + = ith-=<1<= 7.19) « sin'(Z 42) with 2 srs 2 (7.19) where D is the time interval required to traverse the length of the road being affected by the load in any particular location. Differentiating Equation 7.19, substituting it into Equation 7.18, and setting the integration limits from ~D/2 to time ¢ gives: RW) =- | 10 0) sin (EE) a (7.20) ~bp which can be computed numerically. References 1 6 0 R References Burmister, D.M. (1943). The Theory of Stresses and Displacements in Layered Systems and Applications to the Design of Airport Runways, Highway Research Board Proceedings, Washington, DC. Burmister, D.M. (1958). “Evaluation of Pavement Systems of the WASHO Road Test by Layered Systems Methods,” Bulletin 177, Highway Research Board, Washington, DC. Everseries Pavement Analysis Programs (1999). Washington State Department of Transportation, Olympia, WA. Fox, L. (1948). “Computation of Traffic Stresses in a Simple Road Structure,” (1948). Department of Scientific and Industrial Research, Road Research Technical Paper 9, Oxford, UK. Hawng, D., and M.W. Witzack (1979). Program DAMA (Chevron), User’s Manual, Department of Civil Engineering, University of Maryland, College Park, MD. Huang, Y.H. (2004). Pavement Analysis and Design, 2nd ed. Pearson-Prentice Hall, Upper Saddle River, NJ. Kooperman, $., G. Tiller and M. Tseng (1986). ELSYM5, “Interactive Microcomputer Version, User's Manual,” Report FHWA-TS-87-206, Federal Highway Administration, Washington, DC. Odemark N. (1949). Undersokning av Elasticitetegenshapernahos Olika Jordarter samt Teor for Berikning av Beliigningar Enligt Ela sicitetsteorin, Statens Vaginstitut, Meddelande 77 (in Swedish). Ullidtz, P. (1987). Pavement Analysis, Elsevier, Amsterdam, New York. Taylor, D.W. (1963). Fundamentals of Soil Mechanics, John Wiley & Sons, Inc., New York. Timoshenko, 8. P, and J. N. Goodier (1987). Theory of Elasticity, 3rd ed., McGraw-Hill Inc., New York. Tschoegl, N.W. (1989). The Phenomenological Theory of Linear Viscoelastic Behavior: An Introduction, Springer-Verlag, Berlin Hei- delberg. FHWA (1978). VESYS User's Manual: Predictive Design Procedures, FHWA Report FHWA-RD-77-154, Federal Highway Administra- tion, Washington, DG. 203

You might also like