National Academies Press: OpenBook

Rock-Socketed Shafts for Highway Structure Foundations (2006)

Chapter: Chapter Four - Design for Lateral Loading

« Previous: Chapter Three - Design for Axial Loading
Page 53
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 53
Page 54
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 54
Page 55
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 55
Page 56
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 56
Page 57
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 57
Page 58
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 58
Page 59
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 59
Page 60
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 60
Page 61
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 61
Page 62
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 62
Page 63
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 63
Page 64
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 64
Page 65
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 65
Page 66
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 66
Page 67
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 67
Page 68
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 68
Page 69
Suggested Citation:"Chapter Four - Design for Lateral Loading." National Academies of Sciences, Engineering, and Medicine. 2006. Rock-Socketed Shafts for Highway Structure Foundations. Washington, DC: The National Academies Press. doi: 10.17226/13975.
×
Page 69

Below is the uncorrected machine-read text of this chapter, intended to provide our own search engines and external engines with highly rich, chapter-representative searchable text of each book. Because it is UNCORRECTED material, please consider the following text as a useful but insufficient proxy for the authoritative book pages.

SCOPE Twenty-two states indicated in the survey that lateral loading considerations govern the design of drilled shafts in rock or IGM on at least some of their projects. Several states responded that lateral loading governs 100% of their designs. The survey also demonstrated that the most widely used analysis is the p-y method, although other methods are also used. In this chapter, analytical models identified by the literature search and by the survey are re- viewed and evaluated for their applicability to rock sock- ets. Structural issues associated with rock-socketed shafts are considered. DESIGN PROCESS Deep foundations supporting bridge structures may be sub- jected to lateral loading from a variety of sources, including earth pressures, centrifugal forces from moving vehicles, braking forces, wind loading, flowing water, waves, ice, seis- mic forces, and impact. Reese (1984) describes numerous examples of bridges, overhead sign structures, and retaining structures as typical examples of transportation facilities that must sustain significant lateral loading. Drilled shafts are of- ten selected for such structures because they can be designed to sustain lateral loading by proper sizing of the shaft and by providing a sufficient amount of reinforcing steel to resist the resulting bending moments. Design for lateral loading of drilled shafts requires sig- nificant interaction between geotechnical and structural engineers. As described in chapter one, the Bridge Office (structural) is responsible for structural analysis and design of the superstructure and the foundations. However, to model foundation response to lateral loading it is necessary to ac- count for soil/rock-structure interaction. The Geotechnical Division (GD) normally conducts foundation analysis using the models described in this chapter. For preliminary foun- dation design, geotechnical modeling of foundation re- sponse by the GD is used to provide the Bridge Office with information such as depth of fixity, trial designs (diameter and depth) of drilled shafts, and equivalent lateral spring values for use in seismic analysis of the superstructure. The Bridge Office conducts analyses of the superstructure based on models that include fixed-end columns at the depth of fixity. The structural analysis may result in revised loads 54 that are then used by GD to perform revised lateral load analyses. In addition, the Bridge Office may conduct their own analyses using soil and rock-structure interaction mod- els with soil and rock properties provided by GD. The Bridge Office uses the modeling results to establish design parameters for drilled shaft reinforced concrete. According to the 2006 Interim AASHTO LRFD Bridge Design Speci- fications, the strength limit state for lateral resistance of deep foundations is structural only. The basic assumption is that “failure” of the soil/rock does not occur; instead, the geomaterials continue to deform at constant or slightly increasing resistance. Failure occurs when the foundation reaches the structural limit state, defined as the loading at which the nominal combined bending and axial resistance is reached. Axial loading in both compression and uplift requires struc- tural analysis and reinforced-concrete design for drilled shafts. When lateral loading is not significant, structural design is rea- sonably straightforward. When lateral loading is significant, the combined effects of lateral and axial loading are analyzed using models described in this chapter, which account for the effect of axial load by treating the shaft as a beam column. For this reason, structural design issues associated with rock- socketed shafts are addressed in this chapter. ROCK-SOCKETED FOUNDATIONS FOR LATERAL LOADING Rock-socketed shafts provide significant benefits for car- rying lateral loads. Embedment into rock, in most cases, reduces the lateral displacements substantially compared with a deep foundation in soil. To take full advantage of rock-socketed drilled shafts, designers must have confi- dence in the analytical tools used for design. The survey questionnaire shows that traditional methods of analysis for lateral loading of piles and drilled shafts in soil are the most widely used methods currently employed for rock sockets. Recent research has led to some advancements for applying these methods to rock. In addition, several re- searchers have proposed new analytical methods that provide designers with useful tools for predicting load-displacement response and/or structural response of the reinforced- concrete shaft. Each method has advantages and disadvantages for design purposes and these are discussed in the following sections. CHAPTER FOUR DESIGN FOR LATERAL LOADING

55 actual soil/rock reaction. In practice, a great deal of effort and research has been aimed at developing methods for selecting appropriate p-y curves. All of the proposed methods are em- pirical and there is no independent test to determine the rel- evant p-y curve. The principal limitations of the p-y method normally cited are that: (1) theoretically, the interaction of soil or rock between adjacent springs is not taken into ac- count (no continuity), and (2) the p-y curves are not related directly to any measurable material properties of the soil or rock mass or of the foundation. Nevertheless, full-scale load tests and theory have led to recommendations for estab- lishing p-y curves for a variety of soil types (Reese 1984). The method is attractive for design purposes because of the following: • Ability to simulate nonlinear behavior of the soil or rock; • Ability to follow the subsurface stratigraphy (layering) closely; • Can account for the nonlinear flexural rigidity (EI) of a reinforced-concrete shaft; • Incorporates realistic boundary conditions at the top of the foundation; • Solution provides deflection, slope, shear, and moment as functions of depth; • Solution provides information needed for structural de- sign (shear and moment); and • Computer solutions are readily available. Boundary conditions that can be applied at the top of the foundation include: (1) degree of fixity against rotation or translation and (2) applied loads (moment, shear, and axial). With a given set of boundary conditions and a specified family of p-y curves, Eq. 97 is solved numerically, typically using a finite-difference scheme. An iterative solution is required to incorporate the nonlinear p-y curves as well as the nonlinear moment versus EI relationship (material and geometric nonlinearities) for reinforced-concrete shafts. The elastic continuum approach for laterally loaded deep foundations was developed by Poulos (1971), initially for analysis of a single pile under lateral and moment loading at the pile head. The numerical solution is based on the boundary element method, with the pile modeled as a thin elastic strip and the soil modeled as a homogeneous, isotropic elastic ma- terial. This approach was used to approximate socketed piles by Poulos (1972) by considering two boundary conditions at the tip of the pile: (1) the pile is completely fixed against ro- tation and displacement at the tip (rock surface) and (2) the pile is free to rotate but fixed against translation (pinned) at the tip. The fixed pile tip condition was intended to model a socketed deep foundation, whereas the pinned tip was in- tended to model a pile bearing on, but not embedded into, rock. Although these tip conditions do not adequately model the behavior of many rock-socketed shafts, the analyses served to demonstrate some important aspects of socketed deep foundations. For relatively stiff foundations, which FIGURE 36 Subgrade reaction model based on p-y curves (Reese 1997). ANALYTICAL METHODS A laterally loaded deep foundation is the classic example of a soil–structure interaction problem. The soil or rock reaction depends on the foundation displacement, whereas displace- ment is dependent on the soil or rock response and flexural rigidity of the foundation. In most methods of analysis, the foundation is treated as an elastic beam or elastic beam- column. The primary difference in analytical methods used to date lies in the approach used to model the soil and/or rock mass response. Methods of analysis fall into two general cate- gories: (1) subgrade reaction and (2) elastic continuum theory. Subgrade reaction methods treat the deep foundation an- alytically as a beam on elastic foundation. The governing dif- ferential equation (Hetenyi 1946) is given by (97) in which EI = flexural rigidity of the deep foundation, y = lateral deflection of the foundation at a point z along its length, Pz = axial load on the foundation, p = lateral soil/rock reac- tion per unit length of foundation, and w = distributed load along the length of the shaft (if any). In the most commonly used form of subgrade reaction method, the soil reaction–- displacement response is represented by a series of indepen- dent nonlinear springs described in terms of p-y curves (Reese 1984). A model showing the concept is provided in Figure 36. The soil or rock is replaced by a series of discrete mechanisms (nonlinear springs) so that at each depth z the soil or rock reaction p is a nonlinear function of lateral de- flection y. Ideally, each p-y curve would represent the stress–strain and strength behavior of the soil or rock, effects of confining stress, foundation diameter and depth, position of the water table, and any other factors that determine the EI d y dz P d y dz p wz 4 4 2 2 0+ − − =

applies to many drilled shafts, considerable reduction in dis- placement at the pile head can be achieved by socketing, especially if the effect of the socket is to approximate a “fixed” condition at the soil/rock interface. The elastic continuum approach was further developed by Randolph (1981) through use of the finite-element method (FEM). Solutions presented by Randolph cover a wide range of conditions for flexible piles and the results are presented in the form of charts as well as convenient closed-form solutions for a limited range of parameters. The solutions do not ade- quately cover the full range of parameters applicable to rock- socketed shafts used in practice. Extension of this approach by Carter and Kulhawy (1992) to rigid shafts and shafts of intermediate flexibility, as described subsequently, has led to practical analytical tools based on the continuum approach. Sun (1994) applied elastic continuum theory to deep foun- dations using variational calculus to obtain the governing dif- ferential equations of the soil and pile system, based on the Vlasov model for a beam on elastic foundation. This approach was extended to rock-socketed shafts by Zhang et al. (2000), and is also described in this chapter. p-y Method for Rock Sockets The p-y method of analysis, as implemented in various com- puter codes, is the single most widely used method for design of drilled shafts in rock. Responses to the survey ques- tionnaire for this study showed that 28 U.S. state transportation agencies (of 32 responding) use this method. The analytical procedure is dependent on being able to represent the re- sponse of soil and rock by an appropriate family of p-y curves. The only reliable way to verify p-y curves is through instrumented full-scale load tests. The approach that forms the basis for most of the published recommendations for p-y curves in soil is to instrument deep foundations with strain gages to determine the distribution of bending moment over the length of the foundation during a load test. Assuming that the bending moment can be determined reliably from strain gage measurements, the moment as a function of depth can be differentiated twice to obtain p and integrated twice to obtain y. Measured displacements at the foundation head provide a boundary condition at that location. The p-y curves resulting from analysis of field load tests have then been cor- related empirically to soil strength and stress–strain proper- ties determined from laboratory and in situ tests. An alternative approach for deducing p-y curves from load tests is to measure the shape of the deformed foundation; for example, using slope inclinometer measurements and fitting p-y curves to obtain agreement with the measured displace- ments. This approach is described by Brown et al. (1994). Very few lateral load tests on drilled shafts in rock, with the instrumentation necessary to back-calculate p-y curves, 56 have been conducted and published to date. This lack of verification can be viewed as a limitation on use of the p-y method for rock-socketed drilled shafts. A single study by Reese (1997) presents the only published criteria for selec- tion of p-y curves in rock. A few state DOTs have developed in-house correlations for p-y curves in rock. Reese (1997) Reese proposed interim criteria for p-y curves used for analysis of drilled shafts in rock. Reese cautions that the recommendations should be considered as preliminary be- cause of the meager amount of load test data on which they are based. The criteria are summarized as follows. For “weak rock,” defined as rock with unconfined compressive strength between 0.5 MPa and 5 MPa, the shape of the p-y curve, as shown in Figure 37, can be described by the following equations. For the initial linear portion of the curve p = Kiry for y ≤ yA (98) For the transitional, nonlinear portion for y ≥ yA, p ≤ pur (99) yrm = krmB (100) and when the ultimate resistance is reached p = pur (101) where Kir = initial slope of the curve, pur = the rock mass ultimate resistance, B = shaft diameter, and krm is a constant ranging from 0.0005 to 0.00005 that serves to establish the overall stiffness of the curve. p p y y ur rm = ⎛ ⎝⎜ ⎞ ⎠⎟2 0 25. FIGURE 37 Proposed p-y curve for weak rock (Reese 1997).

57 The value of yA corresponding to the upper limit of the initial linear portion of the curve is obtained by setting Eq. 98 equal to Eq. 99, yielding (102) The following expression is recommended for calculating the rock mass ultimate resistance: for 0 ≤ xr ≤ 3B (103) pur = 5.2αrquB for xr ≥ 3B (104) in which qu = uniaxial compressive strength of intact rock, αr = strength reduction factor, and xr = depth below rock surface. Selection of αr is based on the assumption that frac- turing will occur at the surface of the rock under small deflec- tions, thus reducing the rock mass compressive strength. The value of αr is assumed to be one-third for RQD of 100 and to increase linearly to unity at RQD of zero. The underlying as- sumption is that, if the rock mass is already highly fractured, then no additional fracturing with accompanying strength loss will occur. However, this approach appears to have a fundamental shortcoming in that it relies on the compressive strength of the intact rock and not the strength of the rock mass. For a highly fractured rock mass (low RQD) with a high-intact rock strength, it seems that the rock mass strength could be overestimated. The initial slope of the p-y curve, Kir, is related to the ini- tial elastic modulus of the rock mass as follows: (105) where Eir = rock mass initial elastic modulus and kir = di- mensionless constant given by for 0 ≤ xr ≤ 3B (106) for xr > 3B (107) The expressions for kir were determined by fitting a p-y analysis to the results of a field load test (back-fitting) in which the initial rock mass modulus value was determined from PMTs. The method recommended by Reese (1997) is to establish Eir from the initial slope of a pressuremeter curve. Alternatively, Reese suggests the correlation given by Bieni- awski (1978) between rock mass modulus, modulus of intact rock core, and RQD, given as expression 2 in Table 12 (chap- ter two) of this report. According to Reese (1997), rock mass modulus EM determined this way is assumed to be equivalent to Eir in Eq. 105. Results of load tests at two sites are used by Reese (1997) to fit p-y curves according to the criteria given previously for kir = 500 k x Bir r = + ⎛ ⎝⎜ ⎞ ⎠⎟100 400 3 K k Eir ir ir≅ p q B x Bur r u r = +⎛⎝ ⎞⎠α 1 1 4. y p y K A ur rm ir = ( ) ⎡ ⎣ ⎢⎢ ⎤ ⎦ ⎥⎥2 0 25 1 333 . . weak rock. The first load test was located at Islamorada, Florida. A drilled shaft, 1.2 m in diameter and 15.2 m long, was socketed 13.3 m into a brittle, vuggy coral limestone. A layer of sand over the rock was retained by a steel casing and lateral load was applied 3.51 m above the rock surface. The following values were used in the equations for calculating the p-y curves: qu = 3.45 MPa, αr = 1.0, Eir = 7,240 MPa, krm = 0.0005, B = 1.22 m, L = 15.2 m, and EI = 3.73 × 106 kN-m2. Comparison of pile head deflections measured during the load test and from p-y analyses are shown in Figure 38. With a constant value of EI as given above, the analytical results show close agreement with the measured displacements up to a lateral load of about 350 kN. By reducing the values of flexural rigidity in portions of the shaft subject to high mo- ments, the p-y analysis was adjusted to yield deflections that agreed with the measured values at loads higher than 350 kN. The value of krm = 0.0005 was also determined on the basis of establishing agreement between the measured and pre- dicted displacements. The second case analyzed by Reese (1997) is a lateral load test conducted on a drilled shaft socketed into sandstone at a site near San Francisco. The shaft was 2.25 m in diameter with a socket length of 13.8 m. Rock mass strength and mod- ulus values were estimated from PMT results. Three zones of rock were identified and average values of strength and mod- ulus were assigned to each zone. The sandstone is described as “medium-to-fine-grained, well sorted, thinly bedded, very intensely to moderately fractured.” Twenty values of RQD were reported, ranging from zero to 80, with an average of 45. For calculating p-y curves, the strength reduction factor αr was taken as unity, on the assumption that there was “lit- tle chance of brittle fracture.” Values of the other parameters FIGURE 38 Measured and analytical deflection curves for shaft in vuggy limestone (Reese 1997).

used for p-y curve development were: krm = 0.00005; qu = 1.86 MPa for depth of 0–3.9 m, 6.45 MPa for depth of 3.9–8.8 m, and 16.0 MPa for depth of more than 8.8 m; Eir = 10qu (MPa) for each layer, B = 2.25 m, and EI = 35.15 x 103 MN-m2. The value of krm was adjusted to provide agreement between displacements given by the p-y method of analysis and measured displacements from the load test. Figure 39 shows a comparison of the measured load- displacement curve with results produced by the p-y method of analysis, for various methods of computing the flexural rigidity (EI) of the test shaft. Methods that account for the nonlinear relationship between bending moment and EI provide a better fit than p-y analysis with a constant vale of EI. The curve labeled “Analytical” in Figure 39 was ob- tained using an analytical procedure described by Reese to incorporate the nonlinear moment–EI relationships directly into the numerical solution of Eq. 97, whereas the curve la- beled “ACI” incorporates recommendations by the Ameri- can Concrete Institute for treating the nonlinear moment–EI behavior. Fitting of p-y curves to the results of the two load tests as described previously forms the basis for recommendations that have been incorporated into the most widely used com- puter programs being used by state DOTs for analysis of lat- erally loaded rock-socketed foundations. The program COM624 (Wang and Reese 1991) and its commercial version, LPILE (Ensoft, Inc. 2004), allow the user to assign a limited number of soil or rock types to each subsurface layer. One of the options is “weak rock.” If this geomaterial selection is made, additional required input parameters are unit weight, modulus, uniaxial compressive strength, RQD, and krm. The program then generates p-y curves using Eqs. 98–107. The program documentation recommends assigning “weak rock” to geomaterials with uniaxial compressive strengths in the 58 range of 0.5–5 MPa. The user assigns a value to krm. The doc- umentation (Ensoft, Inc. 2004) recommends to: . . . examine the stress–strain curve of the rock sample. Typi- cally, the krm is taken as the strain at 50% of the maximum strength of the core sample. Because limited experimental data are available for weak rock during the derivation of the p-y criteria, the krm from a particular site may not be in the range between 0.0005 and 0.00005. For such cases, you may use the upper bound value (0.0005) to get a larger value of yrm, which in turn will provide a more conservative result. The criteria recommended for p-y curves in the LPILEPLUS users manual (Ensoft Inc. 2004) for “strong rock” is illus- trated in Figure 40. Strong rock is defined by a uniaxial strength of intact rock qu ≥ 6.9 MPa. In Figure 40, su is defined as one-half of qu and b is the shaft diameter. The p-y curve is bilinear, with the break in slope occurring at a deflection y corresponding to 0.04% of the shaft diameter. Resistance (p) is a function of intact rock strength for both portions of the curve. The criterion does not account explic- itly for rock mass properties, which would appear to limit its applicability to massive rock. The authors recommend veri- fication by load testing if deflections exceed 0.04% of the shaft diameter, which would exceed service limit state crite- ria in most practical situations. Brittle fracture of the rock is assumed if the resistance p becomes greater than the shaft diameter times one-half of the uniaxial compressive strength of the rock. The deflection y corresponding to brittle fracture can be determined from the diagram as 0.0024 times the shaft diameter. This level of displacement would be exceeded in many practical situations. It is concluded that the recom- mended criteria applies only for very small lateral deflections and is not valid for jointed rock masses. Some practitioners apply the weak rock criteria, regardless of material strength, to avoid the limitations cited earlier. The authors state that the p-y curve shown in Figure 40 “should be employed with FIGURE 39 Measured and analytical deflection curves, socket in sandstone (Reese 1997). FIGURE 40 Recommended p-y curve for strong rock (Ensoft, Inc. 2004).

59 caution because of the limited amount of experimental data and because of the great variability in rock.” The survey questionnaire for this study found that 28 agencies use either COM624 or LPILEPLUS for analysis of rock-socketed shafts under lateral loading. The following observations are based on a review of the literature: • Existing published criteria for p-y curves in rock are based on a very limited number (two) of full-scale field load tests, • Recommendations for selecting values of input param- eters required by the published criteria are vague and unsubstantiated by broad experience, and • The p-y method of analysis is being used extensively despite these sources of uncertainty. It is therefore concluded that research is needed and should be undertaken with the objective of developing improved criteria for p-y curves in rock. The research should include full-scale field load tests on instrumented shafts, much in the same way that earlier studies focused on the same purpose for deep foun- dations in soil. The p-y curve parameters should be related to rock mass engineering properties that can be determined by state transportation agencies using available site and material characterization methods, as described in chapter two. Current Research by State Agencies The literature review and the survey identified two state trans- portation agencies (North Carolina and Ohio) with research in progress aimed at improving the methodology for construct- ing p-y curves for weathered rock. The North Carolina study is described in a draft report by Gabr et al. (2002) and the Ohio study is summarized in a paper that was under review at the time of this writing, by Liang and Yang (2006). Both studies present recommendations for p-y curves based on a hyperbolic function. Two parameters are required to charac- terize a hyperbola, the initial tangent slope and the asymptote. For the proposed hyperbolic p-y models, these correspond to the subgrade modulus (Kh) and the ultimate resistance (pult), as shown in Figure 41. The hyperbolic p-y relationship is then given as (108) A summary of the two studies, including recommenda- tions for selection of the required parameters (Kh and pult), is presented. In the North Carolina study, results of six full-scale field load tests, at three different sites, were used to develop the p y K y ph ult = + 1 model. Tests were performed on shafts in Piedmont weath- ered profiles of sandstone, mica schist, and crystalline rock. Finite-element modeling was used to calibrate a p-y curve model incorporating subgrade modulus as determined from PMT readings and providing close agreement with strains and deflections measured in the load tests. The model was then used to make forward predictions of lateral load re- sponse for subsequent load tests on socketed shafts at two locations in weathered rock profiles different than those used to develop the model. The procedure for establishing values of subgrade modulus Kh involves determination of the rock mass modulus (EM) from PMT measurements. The coefficient of subgrade reaction is then given by: (109) in which B = shaft diameter, EM = rock mass modulus, vr = Poisson’s ratio of the rock, and Es and Is are modulus and mo- ment of inertia of the shaft, respectively. A procedure is given by Gabr et al. (2002) for establishing the point of rota- tion of the shaft. For p-y curves above the point of rotation, subgrade modulus is equal to the coefficient of subgrade re- action times the shaft diameter or Kh = kh B (110) For depths below the point of rotation, a stiffer lateral sub- grade reaction is assigned and the reader is referred to Gabr et al. (2002) for the equations. An alternative procedure is presented for cases where rock mass modulus is determined using the empirical correlation given by Hoek and Brown (1997) and presented previously as expression 7 in Table 12 of chapter two. In that expression, rock mass modulus is cor- related with GSI and uniaxial compressive strength of intact rock (qu). k E B v E B E I units F lh M r M s s = −( ) ⎡ ⎣⎢ ⎤ ⎦⎥ 0 65 1 2 4 1 12 3 . : ⎛ ⎝⎜ ⎞ ⎠⎟ y p pult k h FIGURE 41 Hyperbolic p-y curve.

The second required hyperbolic curve parameter is the asymptote of the p-y curve, which is the ultimate resistance pult. The proposed expression is given by pult = (pL + τmax)B (111) where pL = limit normal stress and τmax = shearing resistance along the side of the shaft. Gabr et al. adopted the following recommendation of Zhang et al. (2000) for unit side resistance: (112) The limit normal stress is estimated on the basis of Hoek– Brown strength parameters as determined through correla- tions with RMR and GSI, and is given by (113) in which γ ' = effective unit weight of the rock mass, z = depth from the rock mass surface, and the coefficients mb, s, and a are the Hoek–Brown coefficients given by Eqs. 12–15 in chapter two. Results of one of the field load tests conducted for the pur- pose of evaluating the predictive capability of the proposed weak rock (WR) model is shown in Figure 42. The analyses were carried out using the program LPILE. Analyses were also conducted using p-y curves as proposed by Reese (1997), de- scribed previously, as well as several other p-y curve recom- mendations. The proposed model based on hyperbolic p-y curves derived from PMT measurements (labeled dilatometer in Figure 42) shows good agreement with the test results. The authors (Gabr et al. 2002) attributed the underpredicted dis- placements obtained using the Reese criteria to the large values of the factor kir predicted by Eqs. 106 and 107. However, the analysis did not incorporate the nonlinear moment–EI behav- p z q m z q sL u b u a = + + ⎛ ⎝⎜ ⎞ ⎠⎟γ γ ' ' τ max . ( )= 0 20 q MPa u 60 ior of the concrete shaft, which reduces the predicted deflec- tions more significantly than the p-y criteria. One of the lim- itations of the p-y criterion proposed by Gabr et al. (2002) is that it is based on analyses in which EI is taken as a constant. For proper analysis of soil–rock–structure interaction during lateral loading, the nonlinear moment–EI relationship should be modeled correctly. The North Carolina DOT also reports using the program LTBASE, which analyzes the lateral load-displacement response of deep foundations as described by Gabr and Borden (1988). The analysis is based on the p-y method, but also accounts for base resistance by including a vertical resistance component mobilized by shaft rotation and hori- zontal shear resistance, as illustrated in Figure 43. Base re- sistance becomes significant as the relative rigidity of the shaft increases and as the slenderness ratio decreases. For relatively rigid rock sockets, mobilization of vertical and shear resistance at the tip could increase overall lateral ca- pacity significantly, and base resistance effects should be considered. Gabr et al. (2002) stated that the hyperbolic WR p-y curve model is now incorporated into LTBASE, but no results were given. In the Ohio DOT study, Liang and Yang (2006) also pro- pose a hyperbolic p-y curve criterion. The derivation is based on theoretical considerations and finite-element analyses. Results of two full-scale, fully instrumented field load tests are compared with predictions based on the proposed p-y curve criterion. The initial slope of the hyperbolic p-y curve is given by the following semi-empirical equation: (114)K E B B e E I E Bh M ref vr s s M = ⎛ ⎝⎜ ⎞ ⎠⎟ ⎛ ⎝⎜ ⎞ ⎠⎟−2 4 0 284. FIGURE 42 Measured lateral load deflection versus predicted (Gabr et al. 2002). FIGURE 43 Base deformation as a function of shaft rotation (Gabr and Borden 1988).

61 in which Bref = a reference diameter of 0.305 m (1 ft) and all other terms are as defined above for Eq. 109. Liang and Yang (2006) recommend modulus values EM from PMTs for use in Eq. 114, but in the absence of PMT measurements they pre- sent the following correlation equation relating EM to modu- lus of intact rock and GSI: (115) where Er = elastic modulus of intact rock obtained during uniaxal compression testing of core samples. Eq. 115 is also expression 8 of Table 12 in chapter two. Liang and Yang (2006) present two equations for evaluating pult. The first cor- responds to a wedge failure mode, which applies to rock mass near the ground surface. The second applies to rock mass at depth and is given by (116) where pL = limit normal stress, τmax = shearing resistance along the side of the shaft, and pA = horizontal active pressure. Eq. 116 is similar to Eq. 111 (Gabr et al. 2002), but accounts for active earth pressure acting on the shaft. Both methods in- corporate the Hoek–Brown strength criterion for rock mass to evaluate the limit normal stress pL, and both rely on correla- tions with GSI to determine the required Hoek–Brown strength parameters. In the Liang and Yang (2006) approach, pult at each depth is taken as the smaller of the two values obtained from the wedge analysis or by Eq. 116. A source of uncertainty in all of the proposed p-y criteria derives from the choice of method for selecting rock mass modulus when more than one option is available. For exam- ple, using the pressuremeter and GSI data reported by Gabr et al. (2002) significantly different values of modulus are ob- tained for the same site. In some cases, the measured shaft load-displacement response (from load testing) shows better agreement with p-y curves developed from PMT modulus, whereas another load test shows better agreement with p-y curves developed from GSI-derived modulus. Proper selec- tion of rock mass modulus for foundation design is one of the challenges for design of rock-socketed shafts, as pointed out in chapter two. This issue becomes most important when p-y curves for lateral load analysis are based on rock mass modulus. Both the Reese (1997) criteria and the hyperbolic criteria require rock mass modulus to determine the slope of p-y curves. The North Carolina and Ohio programs provide examples of state DOT efforts to advance the state of practice in design of rock-socketed foundations. The programs incorporate careful site investigations using available methods for char- acterizing rock mass engineering properties (RMR, GSI) as well as in situ testing (PMT). Both programs are based on p p p Bult L A= + − ⎛ ⎝⎜ ⎞ ⎠⎟ π τ 4 2 3 max E E eM r= 100 21 7 GSI . analysis of full-scale field load tests on instrumented shafts. However, the proposed equations for generating p-y curves differ between the two proposed criteria and both models will result in different load-displacement curves. It is not clear if either model is applicable to rock sockets other than those used in its development. Both sets of load tests add to the database of documented load tests now available to re- searchers. A useful exercise would be to evaluate the North Carolina proposed criteria against the Ohio load test results and vice versa. Florida Pier Several states reported using other computer programs that are based on the p-y method of analysis. Seven agencies re- port using the Florida Bridge Pier Analysis Program (FBPIER) for analysis of rock-socketed shafts. Of those seven, six also report using COM624 and/or LPILE. The FBPIER, described by Hoit et al. (1997), is a nonlinear, finite- element analysis program designed for analyzing bridge pier substructures composed of pier columns and a pier cap sup- ported on a pile cap and piles or shafts including the soil (or rock). FBPIER was developed to provide an analytical tool allowing the entire pier structure of a bridge to be analyzed at one time, instead of multiple iterations between foundation analysis programs (e.g., COM624) and structural analysis programs. Basically, the structural elements (pier column, cap, pile cap, and piles) are modeled using standard structural finite-element analysis, including nonlinear capabilities (nonlinear M–EI behavior), whereas the soil response is modeled by nonlinear springs (Figure 44). Axial soil re- sponse is modeled in terms of t-z curves, whereas lateral response is modeled in terms of p-y curves. The program has built-in criteria for p-y curves in soil, based on published recommendations and essentially similar to those employed in LPILE. User-defined p-y curves can also be specified. To simulate rock, users currently apply the criteria for either soft clay (Matlock 1970) or stiff clay (Reese and Welch 1975) but with strength and stiffness properties of the rock, or user- defined curves are input. Research is underway to incorpo- rate improved p-y curve criteria into FBPIER, specifically for FIGURE 44 Florida pier model for structure and foundation elements (Hoit et al. 1997).

Florida limestone, as described by McVay and Niraula (2004). Centrifuge tests were conducted in which instrumented model shafts embedded in a synthetic rock (to simulate Florida limestone) were subjected to lateral loading. Strain gage measurements were used to back-calculate p-y curves, which are presented in normalized form, with p normalized by shaft diameter and rock compressive strength (p/Bqu) and y normalized by shaft diameter (y/B). There is no analytical expression recommended for new p-y curve criteria and the report recommends that field testing be undertaken on full- size drilled shafts to validate the derived p-y curves estab- lished from the centrifuge tests before they are employed in practice. Strain Wedge Model The strain wedge (SW) model has been applied to laterally loaded piles in soil, as described by Ashour et al. (1998). The 2006 Interim AASHTO LRFD Bridge Design Specifications identify the SW model as an acceptable method for lateral load analysis of deep foundations. The 3-D soil–pile interac- tion behavior is modeled by considering the lateral resistance that develops in front of a mobilized passive wedge of soil at each depth. Based on the soil stress–strain and strength prop- erties, as determined from laboratory triaxial tests, the hori- zontal soil strain (ε) in the developing passive wedge in front of the pile is related to the deflection pattern (y) versus depth. The horizontal stress change (ΔσH) in the developing passive wedge is related to the soil–pile reaction (p), and the nonlin- ear soil modulus is related to the nonlinear modulus of sub- grade reaction, which is the slope of the p-y curve. The SW model can be used to develop p-y curves for soil that show good agreement with load test results (Ashour and Norris 2000). Theoretically, the SW model overcomes some of the limitations of strictly empirically derived p-y curves because the soil reaction (p) at any given depth depends on the re- sponse of the neighboring soil layers (continuity) and prop- erties of the pile (shape, stiffness, and head conditions). Ashour et al. (2001) proposed new criteria for p-y curves in weathered rock for use with the SW model. The criteria are described by the authors as being based on the weak rock criteria of Reese (1997) as given by Eqs. 98-104, but modi- fied to account for the nonlinear rock mass modulus and the strength of the rock mass in terms of Mohr–Coulomb strength parameters c and φ. Ashour et al. (2001) reported good agreement between the SW analysis and a field load test reported by Brown (1994). One state DOT (Washington) reports using the computer program (S-Shaft) based on the SW model that incorporates the p-y curve criteria for weath- ered rock. However, the program has not yet been used for design of a socketed shaft (J. Cuthbertson, personal com- munication, Sep. 30, 2005). The SW model and proposed p-y criteria of Ashour et al. (2001) warrant further consider- ation and should be evaluated against additional field load test results (e.g., the tests reported by Gabr et al. 2002 and Liang and Yang 2006). 62 Continuum Models for Laterally Loaded Sockets Carter and Kulhawy (1992) Carter and Kulhawy (1988, 1992) studied the behavior of flex- ible and rigid shafts socketed into rock and subjected to lateral loads and moments. Solutions for the load-displacement rela- tions were first generated using finite-element analyses. The finite-element analyses followed the approach of Randolph (1981) for flexible piles under lateral loading. Based on the FEM solutions, approximate closed-form equations were developed to describe the response for a range of rock-socket parameters typically encountered in practice. The results pro- vide a first-order approximation of horizontal groundline dis- placements and rotations and can incorporate an overlying soil layer. The method is summarized as follows. Initially, consider the case where the top of the shaft cor- responds to the top of the rock layer (Figure 45). The shaft is idealized as a cylindrical elastic inclusion with an effective Young’s modulus (Ee), Poisson’s ratio (vc), depth (D), and di- ameter (B), subjected to a known lateral force (H), and an overturning moment (M). For a reinforced-concrete shaft having an actual flexural rigidity equal to (EI)c, the effective Young’s modulus is given by (117) It is assumed that the elastic shaft is embedded in a ho- mogeneous, isotropic elastic rock mass, with properties Er and vr. Effects of variations in the Poisson’s ratio of the rock mass (vr), are represented approximately by an equivalent shear modulus of the rock mass (G*), defined as: (118) in which Gr = shear modulus of the elastic rock mass. For an isotropic rock mass, the shear modulus is related to Er and vr by G G vr r∗ = + ⎛ ⎝⎜ ⎞ ⎠⎟1 3 4 E EI Be c = ( ) π 4 64 FIGURE 45 Lateral loading of rock-socketed shaft (Carter and Kulhawy 1992).

63 (119) Based on a parametric study using finite-element analysis, it was found that closed-form expressions could be obtained to provide reasonably accurate predictions of horizontal dis- placement (u) and rotation (θ) at the head of the shaft for two limiting cases. The two cases correspond to flexible shafts and rigid shafts. The criterion for a flexible shaft is (120) For shafts satisfying Eq. 120, the response depends only on the modulus ratio (Ee/G*) and Poisson’s ratio of the rock mass (vr) and is effectively independent of D/B. The follow- ing closed-form expressions, suggested by Randolph (1981), provide accurate approximations for the deformations of flexible shafts: (121) (122) in which u = groundline deflection and θ = groundline rota- tion of the shaft. Carter and Kulhawy (1992) reported that the accuracy of the above equations is verified for the following ranges of pa- rameters: 1 ≤ Ee/Er ≤ 106 and D/B ≥ 1. The criterion for a rigid shaft is (123) and (124) When Eqs. 123 and 124 are satisfied, the displacements of the shaft will be independent of the modulus ratio (Ee/Er) and will depend only on the slenderness ratio (D/B) and Poisson’s ratio of the rock mass (vr). The following closed-form expres- sions give reasonably accurate displacements for rigid shafts: (125) (126)θ = ⎛⎝⎜ ⎞⎠⎟ ⎛⎝⎜ ⎞⎠⎟ + ⎛⎝⎜ ⎞⎠⎟∗ − ∗ 0 3 2 0 82 7 8 3. . H G B D B M G B 2 5 3D B ⎛⎝⎜ ⎞⎠⎟ − u H G B D B M G B D B = ⎛⎝ ⎞⎠ ⎛⎝ ⎞⎠ + ⎛⎝ ⎞⎠ ⎛⎝ ⎞⎠∗ − ∗ 0 4 2 0 3 2 1 3 2 . . − 7 8 E G B D e ∗ ( ) ≥2 1002 D B E G e≤ ⎛⎝⎜ ⎞ ⎠⎟∗0 05 1 2 . θ = ⎛⎝⎜ ⎞ ⎠⎟ ⎛ ⎝⎜ ⎞ ⎠⎟ + ⎛ ⎝⎜∗ ∗ − ∗ 1 08 6 40 2 3 7 3. . H G B E G M G B e ⎞⎠⎟ ⎛ ⎝⎜ ⎞ ⎠⎟∗ −E G e 5 7 u H G B E G M G B e = ⎛ ⎝⎜ ⎞ ⎠⎟ ⎛ ⎝⎜ ⎞ ⎠⎟ + ⎛ ⎝⎜ ⎞ ∗ ∗ − ∗ 0 50 1 08 1 7 2 . . ⎠⎟ ⎛ ⎝⎜ ⎞ ⎠⎟∗ −E G e 3 7 D B E G e≥ ⎛⎝⎜ ⎞ ⎠⎟∗ 2 7/ G E v r r r = +( )2 1 The accuracy of Eqs. 125 and 126 has been verified for the following ranges of parameters: 1 ≤ D/B ≤ 10 and Ee/Er ≥ 1. Shafts can be described as having intermediate stiffness whenever the slenderness ratio is bounded approximately as follows: (127) For the intermediate case, Carter and Kulhawy suggested that the displacements be taken as 1.25 times the maximum of either (1) the predicted displacement of a rigid shaft with the same slenderness ratio (D/B) as the actual shaft or (2) the predicted displacement of a flexible shaft with the same mod- ulus ratio (Ee/G*) as the actual shaft. Values calculated in this way should, in most cases, be slightly larger than those given by the more rigorous finite-element analysis for a shaft of in- termediate stiffness. Carter and Kulhawy next considered a layer of soil of thickness Ds overlying rock, as shown in Figure 46. The analysis is approached by structural decomposition of the shaft and its loading, as shown in Figure 46b. It was assumed that the magnitude of applied lateral loading is sufficient to cause yielding within the soil from the ground surface to the top of the rock mass. The portion of the shaft within the soil is then analyzed as a determinant beam subjected to known loading. The displacement and rotation of point A relative to point O can be determined by established techniques of struc- tural analysis. The horizontal shear force (Ho) and bending moment (Mo) acting in the shaft at the rock surface level can be computed from statics, and the displacement and rotation at this level can be computed by the methods described pre- viously. The overall groundline displacements can then be calculated by superposition of the appropriate parts. Determination of the limiting soil reactions is recom- mended for the two limiting cases of cohesive soil in undrained loading (φ = 0) and frictional soil (c = 0) in drained loading. Ultimate resistance for shafts in cohesive soils is based on the method of Broms (1964a), in which the undrained 0 05 1 2 2 7 . E G D B E G e e ∗ ∗ ⎛ ⎝⎜ ⎞ ⎠⎟ < < ⎛ ⎝⎜ ⎞ ⎠⎟ FIGURE 46 Rock-socketed shaft with overlying layer (Carter and Kulhawy 1992).

soil resistance ranges from zero at the ground surface to a depth of 1.5B and has a constant value of 9su below this depth, where su = soil undrained shear strength. For socketed shafts extending through a cohesionless soil layer, the following lim- iting pressure suggested by Broms (1964b) is assumed: (128) (129) in which σv' = vertical effective stress and φ' = effective stress friction angle of the soil. For both cases (undrained and drained) solutions are given by Carter and Kulhawy (1992) for the displacement, rotation, shear, and moment at point O of Figure 46. The contribution to groundline displacement and rotation from the loading transmitted to the rock mass (Ho and Mo) is determined based on Eqs. 121 and 122 or Eqs. 125 and 126 and added to the calculated displacement and rotation at the top of the socket to determine overall groundline response. Application of the proposed theory is described by Carter and Kulhawy (1992) through back-analysis of a single case in- volving field loading of a pair of rock-socketed shafts. The method has not been evaluated against a sufficient database of field performance, and further research is needed to assess its reliability. The analysis was developed primarily for ap- plication to electrical transmission line foundations in rock, although the concepts are not limited to foundations support- ing a specific type of structure. The approach is attractive for design purposes, because the closed-form equations can be executed by hand or on a spreadsheet. Carter and Kulhawy (1992) stated that the assumption of yield everywhere in the soil layer may represent an oversim- plification, but that the resulting predictions of groundline displacements will overestimate the true displacements, giv- ing a conservative approximation. However, the assumption that the limit soil reaction is always fully mobilized may lead to erroneous results by overestimating the load carried by the soil and thus underestimating the load transmitted to the socket. Furthermore, groundline displacements may be un- derestimated because actual soil resistance may be smaller than the limiting values assumed in the analysis. Zhang et al. (2000) Zhang et al. (2000) extended the continuum approach to pre- dict the nonlinear lateral load-displacement response of rock- socketed shafts. The method considers subsurface profiles consisting of a soil layer overlying a rock layer. The defor- mation modulus of the soil is assumed to vary linearly with depth, whereas the deformation modulus of the rock mass is assumed to vary linearly with depth and then to stay constant below the shaft tip. Effects of soil and/or rock mass yielding on response of the shaft are considered by assuming that the K p = + − 1 1 sin ' sin ' φ φ p Ku p v= 3 σ ' 64 soil and/or rock mass behaves linearly elastically at small strain levels and yields when the soil and/or rock mass reac- tion force p (force/length) exceeds the ultimate resistance pult (force/length). Analysis of the loaded shaft as an elastic continuum is ac- complished using the method developed by Sun (1994). The numerical solution is by a finite-difference scheme and in- corporates the linear variation in soil modulus and linear variation in rock mass modulus above the base of the shaft. Solutions obtained for purely elastic responses are compared with those of Poulos (1971) and finite-element solutions by Verruijt and Kooijman (1989) and Randolph (1981). Rea- sonable agreement with those published solutions is offered as verification of the theory, for elastic response. The method is extended to nonlinear response by account- ing for local yielding of the soil and rock mass. The soil and rock mass are modeled as elastic, perfectly plastic materials, and the analysis consists of the following steps: 1. For the applied lateral load H and moment M, the shaft is analyzed by assuming the soil and rock mass are elastic, and the lateral reaction force p of the soil and rock mass along the shaft is determined by solution of the governing differential equation and boundary con- ditions at the head of the shaft. 2. The computed lateral reaction force p is compared with the ultimate resistance pult. If p > pult, the depth of yield zy in the soil and/or rock mass is determined. 3. The portion of the shaft in the unyielded soil and/or rock mass (zy ≤ z ≤ L) is considered to be a new shaft and analyzed by ignoring the effect of the soil and/or rock mass above the level z = zy. The lateral load and moment at the new shaft head are given by: (130) (131) 4. Steps 2 and 3 are repeated and the iteration is continued until no further yielding of soil or rock mass occurs. 5. The final results are obtained by decomposition of the shaft into two parts, which are analyzed sepa- rately, as illustrated previously in Figure 46. The sec- tion of the shaft in the zone of yielded soil and/or rock mass is analyzed as a beam subjected to a dis- tributed load of magnitude pult. The length of shaft in the unyielded zone of soil and/or rock mass is ana- lyzed as a shaft with the soil and/or rock mass behaving elastically. Ultimate resistance developed in the overlying soil layer is evaluated for the two conditions of undrained loading (φ = 0) and fully drained loading (c = 0). For fine-grained soils (clay), undrained loading conditions are assumed and the limit pressure is given by M M Hz p z z dzo y ult zy y= + − −( )∫0 H H p dzo ult zy = − ∫0

65 (132) (133) in which cu = undrained shear strength, B = shaft diameter, γ ' = average effective unit weight of soil above depth z, J = a coef- ficient ranging from 0.25 to 0.5, and R = shaft radius. For shafts in sand, a method attributed to Fleming et al. (1992) is given as follows: (134) where Kp = Rankine coefficient of passive earth pressure de- fined by Eq. 129. Ultimate resistance of the rock mass is given by (135) where τmax = maximum shearing resistance along the sides of the shaft (e.g., Eq. 30 of chapter three) and pL = normal limit resistance. The limit normal stress pL is evaluated using the Hoek–Brown strength criterion with the strength parameters determined on the basis of correlations to GSI. The resulting expression was given previously as Eq. 113. According to Zhang et al. (2000), a computer program was written to execute this procedure. Predictions using the pro- posed method are compared with results of field load tests re- ported by Frantzen and Stratten (1987) for shafts socketed into sandy shale and sandstone. Computed pile head deflections show reasonable agreement with the load test results. The method appears to have potential as a useful tool for founda- tions designers. Availability of the computer program is un- known. Programming the method using a finite-difference scheme as described by Zhang et al. (2000) is also possible. Discontinuum Models A potential mode of failure for a laterally loaded shaft in rock is by shear failure along joint surfaces. To et al. (2003) pro- posed a method to evaluate the ultimate lateral-load capacity of shafts in rock masses with two or three sets of intersecting joints. The analysis consists of two parts. In the first part, the block theory of Goodman and Shi (1985) is used to deter- mine if possible combinations of removable blocks exist that would represent a kinematically feasible mode of failure. In the second part, the stability of potentially removable combinations of blocks or wedges is analyzed by limit equi- librium. Both steps in the analysis require careful evaluation of the joint sets, in terms of their geometry and strength prop- erties. Although the method is based on some idealized assumptions, such as equal joint spacing, and it has not been evaluated against field or laboratory load tests, it provides a theoretically based discontinuum analysis of stability in cases where this mode of failure requires evaluation. p p Bult L= +( )τmax p K zBult p= 2γ ' N c z J R zp u = + + ≤3 2 9γ ' p N c Bult p u= Discussion of Analytical Models for Laterally Loaded Sockets Each of the analytical methods described above has advan- tages and disadvantages for use in the design of rock-socketed shafts for highway bridge structures. The greatest need for further development of all available methods is a more thor- ough database of load test results against which existing theory can be evaluated, modified, and calibrated. The simple closed-form expressions given by Carter and Kulhawy (1992) represent a convenient, first-order approxi- mation of displacements and rotations of rock-socketed shafts. Advantages include the following: • Predicts lateral displacements under working load conditions, • Requires a single material parameter (rock mass modulus), • Provides reasonable agreement with theoretically rigorous finite-element analysis, and • Is the easiest method to apply by practicing design engineers. Limitations include: • Does not predict the complete lateral load-displacement curve, • Elastic solution does not provide shear and moment dis- tribution for structural design, • Does not account for more than one rock mass layer, • Does not account directly for nonlinear M–EI behavior of reinforced-concrete shaft, and • Does not account for interaction between axial and lateral loading and its effects on structural behavior of the shaft. The method can be best used for preliminary design; for example, establishing the initial trial depth and diameter of rock-socketed shafts under lateral and moment loading. For some situations, no further analysis may be necessary. Final design should be verified by field load testing. The method of Zhang et al. (2000) provides a more rigorous continuum-based analysis than that of Carter and Kulhawy. The tradeoff is that more material parameters are required as input. Variation of rock mass modulus with depth is required. To fully utilize the nonlinear capabilities, the Hoek–Brown yield criterion parameters are required, and these are based on establishing the RMR and/or GSI. The method is best applied when a more refined analysis is required and the agency is will- ing to invest in proper determination of the required material properties. Advantages include: • Predicts the full, nonlinear, lateral load-deformation re- sponse; • Accounts for partial yield in either the rock mass or the overlying soil (more realistic);

• Is based on well-established rock mass and soil properties; • Is verified against rigorous theory, for elastic range; and • Provides shear and moment distribution for structural design. Limitations include: • Requires numerical (computer) solution, not currently available commercially; • Requires a larger number of rock mass material para- meters; • Currently is limited to two layers (one soil and one rock mass layer, or two rock mass layers); and • Nonlinear M–EI behavior of reinforced-concrete shaft is not accounted for explicitly; requires iterative analy- ses with modified values of EI. The most rigorous analytical methods based on a continuum approach are FEM. When implemented by competent users, FEM analysis can account for the shaft, soil, and rock mass behaviors more rigorously than the approximate methods de- scribed herein, but FEM analyses are not suitable for routine design of foundations in most cases. First, the results are only as reliable as the input parameters. In most cases the material properties of the rock mass are not known with sufficient reli- ability to warrant the more sophisticated analysis. Second, the design engineer should have the appropriate level of knowledge of the mathematical techniques incorporated into the FEM analyses. Finally, the time, effort, and expense required for conducting FEM analyses are often not warranted. For very large or critical bridge structures, sophisticated FEMs may be warranted and the agency might benefit from the investment required in computer codes, personnel training, and field and laboratory testing needed to take advantage of such techniques. Subgrade reaction methods, as implemented through the p-y curve method of analysis, offer some practical advantages for design. These include: • Predicts the full, nonlinear lateral load-deformation response; • Can incorporate multiple layers of soil and/or rock; • Accounts for nonlinear M–EI behavior of reinforced- concrete shaft; • Provides structural analysis (shear, moment, rotation, and displacement) of the drilled shaft; • Accounts for the effects of axial compression load on the structural behavior of the shaft; and • Can be implemented easily on a desktop computer with available software. The principal limitations are: • Lack of a strong theoretical basis for p-y curves and • Requires back analysis of instrumented load tests to verify and validate p-y curves; such verification is cur- rently lacking or limited to a few cases. 66 Considering that the p-y method is currently being used extensively by most state DOTs, effort should be made to address its present limitation by research aimed at better es- tablishing methods to specify appropriate p-y curves in rock. Full-scale field load testing with instrumentation is the only known method to verify p-y curves. Research conducted for this purpose would provide an opportunity to evaluate and calibrate other proposed analytical methods; for example, those of Carter and Kulhawy (1992) and Zhang et al. (2000) and for development of new models. Recommendations for research are discussed further in chapter five. The research programs sponsored by the North Carolina and Ohio DOTs illustrate the type of approach that is useful for advancing all of the available methods of analysis. In addition to providing improved criteria for p-y curve modeling, the load test results reported by Gabr et al. (2002) and Liang and Yang (2006) can be used to evaluate each other’s models and the SW and continuum models described in this chapter. In summary, a range of analytical tools are available to foundation designers to consider rock sockets under lateral and moment loading. These include simple, closed-form equations requiring a small number of material properties (Carter and Kulhawy 1992). A more rigorous model that predicts the complete nonlinear response but requires more material properties is also available (Zhang et al. 2000). Highly sophisticated numerical models requiring extensive material properties and appropriate expertise (FEM analysis) exist and may be appropriate for larger projects. The p-y method of analysis is attractive to designers, as evidenced by its wide use; however, considerable judgment is required in selection of p-y curve parameters. All of the currently avail- able methods suffer from a lack of field data for verification and are best applied in conjunction with local and agency ex- perience, thorough knowledge of the geologic environment, and field load testing. STRUCTURAL ISSUES Twenty of the questionnaire responses indicated that struc- tural design of drilled shaft foundations is carried out by en- gineers in the Bridge Design or Structures Division of their state DOTs. Three states indicated that structural design is a joint effort between the Geotechnical and Structural/Bridge Divisions. One DOT indicated that structural design is done by the Geotechnical Branch. All of the states responding to the structural design portion of the questionnaire stated that the AASHTO LRFD Bridge Design Specifications are fol- lowed for structural design of drilled shafts. Three states also cited the ACI Building Code Requirements for Structural Concrete. Barker et al. (1991) discussed the structural design of reinforced-concrete shafts and have several design examples illustrating the basic concepts. O’Neill and Reese (1999) also covered the general aspects of reinforced-concrete design for

67 drilled shafts, for axial compression loading and flexure, citing as primary references the 1994 AASHTO LRFD Bridge Design Specifications (1st edition) and the 1995 ACI Building Code Requirements for Structural Concrete (ACI 318-94). Both the AASHTO and ACI codes have since been revised (AASHTO in 2004 and ACI in 2002); however, there are no major differences that would change the structural design of drilled shafts. According to the survey, all of the states are designing in accordance with the AASHTO LRFD Bridge Design Specifications. At the time of this writing, Section 10 (Foundations) of the draft 2006 Interim AASHTO LRFD Bridge Design Specifications was available for reference. However, the other sections of the 2006 Interim specifica- tions were not available and so comments pertaining to Sec- tion 5 are referenced to the 2004 edition. Only issues of struc- tural design pertaining specifically to rock-socketed drilled shafts are addressed here. General Issues Section 10.8.3.9 (“Shaft Structural Resistance”) of the 2006 Interim AASHTO LRFD Bridge Design Specifications states that The structural design of drilled shafts shall be in accordance with the provisions of Section 5 for the design of reinforced concrete. This language makes it clear that drilled shaft structural design is subject to the same provisions as other reinforced- concrete members. The designer must then determine whether the shaft is a compression member or a member sub- jected to compression and flexure (beam column). Article 10.8.3.9.3 states the following: Where the potential for lateral loading is insignificant, drilled shafts may be reinforced for axial loads only. Those portions of drilled shafts that are not supported laterally shall be designed as reinforced-concrete columns in accordance with Article 5.7.4. Reinforcing steel shall extend a minimum of 10 ft below the plane where the soil provides fixity. The commentary accompanying Article 10.8.3.9.3 states fur- ther that: A shaft may be considered laterally supported: below the zone of liquefaction or seismic loads, in rock, or 5.0 ft below the ground surface or the lowest anticipated scour elevation. . . . . Laterally supported does not mean fixed. Fixity would occur somewhere below this location and depends on the stiffness of the supporting soil. The language in this provision could be improved by pro- viding a definition of “fixity.” Fixity is defined by Davisson (1970) for piles under lateral loading as the depth below groundline corresponding to the fixed base of an equivalent free-standing column; that is, a column for which the top deflection and rotation would be the same as that of a column supported by the embedded deep foundation (Figure 47). Approximate equations are given by Davisson for establishing depth of fixity based on the depth of the foundation and a relative stiffness factor that depends on the flexural rigidity of the pile and the subgrade modulus of the soil or rock. In- terviews with state DOT engineers indicated that different criteria for establishing depth of fixity are being applied. One state DOT defines fixity as the depth at which LPILE analy- sis shows the maximum moment, whereas another defines fixity as the depth at which LPILE shows zero lateral deflec- tion. In Section 12 of Bridge Design Aids (1990), the Massa- chusetts Highway Department) describes a rigorous ap- proach involving use of the program LPILE (or other p-y analysis) to establish a depth of fixity as defined in Figure 47. For the given soil/rock profile, approximate service loads are applied to the “Top of Column” (Figure 47). Shear and moment are applied as separate load cases and the resulting lateral deflections and rotations at the top of the column are designated as follows: δV = deflection due to shear (V) δM = deflection due to moment (M) θV = rotation due to shear (V) θM = rotation due to moment (M). Equivalent column lengths are then calculated using the following analytical expressions for each loading case. The four resulting values of L should be approximately equal and the average value can be taken as a reasonable approxima- tion of the equivalent fixed-end column length. Depth of fix- ity corresponds to the portion of the fixed-end column below groundline. (136) (137)L EI MM M δ δ = ( )⎡ ⎣⎢ ⎤ ⎦⎥ 2 1 2 L EI VV V δ δ = ( )⎡ ⎣⎢ ⎤ ⎦⎥ 3 1 3 PP V M M depth of fixity y y Deep Foundation Top of Column V Equivalent Fixed-end Column FIGURE 47 Depth of fixity for equivalent fixed-end column.

(138) (139) The principal use of depth of fixity is to establish the elevation of equivalent fixed-end columns supporting the superstructure, thus enabling structural designers to uncou- ple the foundations from the superstructure for the purpose of structural analysis and design of the bridge or other structure. Structural modeling of the superstructure with equivalent fixed-end columns is also used to establish the column loads. These column loads are then used to analyze the drilled shaft foundations by applying them to the top of the actual col- umn, which is continuous with the foundation (left side of Figure 47) using p-y analysis. As described at the beginning of this chapter, these analyses may be done by either the GD or Bridge offices, but the soil and rock parameters are pro- vided by GD. The p-y analysis gives the maximum moment and shear that are used in the reinforced-concrete design. Use of software such as LPILE, COM624, or other programs is thus seen to be an integral tool in both the geotechnical and structural design of drilled shafts for bridges or other transportation structures. As noted previously, AASHTO specifications define the strength limit state for lateral load- ing only in terms of foundation structural resistance. Lateral deflections as predicted by p-y analyses are used as a design tool to satisfy service limit state criteria. The concept of fixity also has implications for reinforcing steel requirements of drilled shafts. According to Article 10.8.3.9.3, as cited earlier, if a drilled shaft designed for axial compression extends through soil for a distance of at least 3 m (10 ft) beyond fixity before entering into rock, the rock- socketed portion of the shaft does not require reinforcement. This provision would also limit the need for compression steel in rock sockets to a maximum depth of 3 m below fix- ity. Exceptions to this are shafts in Seismic Zones 3 and 4, for which Article 5.13.4.6.3d states that “for cast-in-place piles, longitudinal steel shall be provided for the full length of the pile.” Some state DOTs use permanent steel casing in the top por- tion of drilled shafts or, in many cases, down to the top of rock. Permanent casing is not mentioned in the 2006 Interim speci- fications, but the 2004 specifications included the following statement: “Where permanent steel casing is used and the shell is smooth pipe greater than 0.12 in. thick, it may be considered to be load-carrying. Allowance should be made for corrosion.” A few states indicated that questions arise in connection with relatively short sockets in very hard rock. The questions pertain to moment transfer, development length of steel re- inforcing, and apparently high shear loads resulting from high moment loading. L EI MM M θ θ = ( ) L EI VV V θ θ = ( )⎡ ⎣⎢ ⎤ ⎦⎥ 2 1 2 68 Moment Transfer Rock sockets subjected to high lateral and/or moment loading require a minimum depth of embedment to transfer moment to the rock mass and to satisfy minimum development length requirements for reinforcing steel. The mechanism of moment transfer from a column to the rock is through the lateral resistance developed between the concrete shaft and the rock. The resistance depends on many of the factors identified previously, primarily strength and stiffness of the rock mass and flexural rigidity of the shaft. When the strength and mod- ulus of the rock mass are greater than that of the concrete shaft, the question may arise, why excavate such high qual- ity rock and replace it with lower strength concrete? The only means to transfer moment into the rock mass is through a properly designed shaft with the dimensions, strength, and stiffness to transmit the design moment by the assumed mechanisms of lateral resistance. In some situations where high strength rock mass is close to the ground surface, shaft size may be governed by structural considerations rather than by geotechnical capacity. For some relatively short, stubby shafts in hard rock, socket length could be governed by the required develop- ment length of longitudinal reinforcing bars. Article 5.11 of the AASHTO LRFD Bridge Design Specifications (2004) specifies basic tension development lengths for various bar sizes as a function of steel and concrete strengths. Because the bars will be stressed to their maximum values at the points where maximum moments occur, the distance between the point of maximum moment and the bottom of the socket must be at least equal to the required development length. As an example, for No. 18 bars, assuming fy = 414 MPa (60 ksi) and fc' = 27.6 MPa (4 ksi), basic development length is 267 cm (105 in. or 8.75 ft). Although this is not often the governing factor for socket length, it should be checked. Shear Some designers commented on cases where p-y analysis of laterally loaded rock-socketed shafts resulted in unexpectedly high values of shear and whether the results were realistic. In particular, when a rock socket in relatively strong rock is subjected to a lateral load and moment at its head, values of shear near the top of the socket may be much higher than the applied lateral load. This result would be expected given the mechanism of moment load transfer. When the lateral load has a high moment arm, such as occurs in an elevated struc- ture, the lateral load transmitted to the top of the drilled shaft may be small or modest, but the moment may be relatively large. The principal mechanism of moment transfer from the shaft to the rock mass is through the mobilized lateral resis- tance. If a large moment is transferred over a relatively short depth, the lateral resistance is also concentrated over a rela- tively short length of the shaft and results in shear loading that may be higher in magnitude than that of the lateral load.

69 There is some question, however, whether high values of shear predicted by p-y methods of analysis for such cases exist in reality or are artifacts of the analysis. One designer sug- gested that the structural model of the shaft does not account properly for shear deformation, resulting in unrealistically high shear values. The topic requires further investigation. In some cases, the magnitude of shear must be addressed in the reinforced-concrete design, primarily in the use of transverse reinforcement. According to AASHTO LRFD Bridge Design Specifications (2004), the minimum amount of spiral reinforcement to satisfy the requirements for com- pression is governed by (140) in which ρs = ratio of spiral reinforcement to total volume of concrete core, measured out-to-out of spirals; Ag = gross (nominal) cross-sectional area of concrete; and Ac = cross- sectional area of concrete inside the spiral steel. When shear occurs in addition to axial compression, the section is then checked by comparing the factored shear loading with the factored shear resistance, given by (141) in which Vr = factored shear resistance, Vn = nominal shear resistance, Vc = nominal shear resistance provided by the concrete, Vs = nominal shear resistance provided by the trans- verse steel, and φ = resistance factor = 0.90 for shear. The nominal shear strength provided by the concrete is given (in U.S. customary units) by: (142) or, when axial load is zero, (143) where Pu = factored axial load and Av = area of concrete in the cross section that is effective in resisting shear. For a circular section this can be taken as (144) in which B = shaft diameter and Bls = diameter of a circle passing through the center of the longitudinal reinforcement. The nominal shear strength provided by transverse rein- forcement is given by (145)V A f d s s vs y = A B B Bv ls= + ⎡ ⎣⎢ ⎤ ⎦⎥0 9 2. π V f Ac c v= 2 ' V P A f Ac u g c v= + ⎛ ⎝⎜ ⎞ ⎠⎟2 1 2 000, ' V V V Vr n c s= = +( )φ φ ρs g c c y A A f f= − ⎛ ⎝⎜ ⎞ ⎠⎟0 45 1. ' where Avs = area of shear reinforcement, s = longitudinal (ver- tical) spacing of the ties or pitch of the spiral, and d = effective shear depth. For a circular cross section this can be taken as (146) The need for additional transverse reinforcement, beyond that required for compression, can be determined by Eq. 141. For the majority of rock-socketed shafts, the transverse rein- forcement required to satisfy compression criteria (Eq. 140) combined with the shear resistance provided by the concrete (Eq. 142) will be adequate to resist the factored shear load- ing without the need for additional transverse reinforcement. However, in cases where high lateral load or moment are to be distributed to the ground over a relatively small distance; for example, a short stubby socket in high-strength rock, factored shear forces may be high and the shaft dimensions and reinforcement may be governed by shear. In these cases, the designer is challenged to provide a design that provides adequate shear resistance without increasing the costs exces- sively or adversely affecting constructability by constricting the flow of concrete. To handle high shear loading in the reinforced-concrete shaft, the designer has several options: (1) increase the shaft diameter, thus increasing the area of shear-resisting concrete; (2) increase the shear strength of the concrete; or (3) increase the amount of transverse reinforcing, either spiral or ties, to carry the additional shear. Each option has advantages and disadvantages. Two variables that can be adjusted to increase shear re- sistance are concrete 28-day compressive strength, f 'c, and shaft diameter, B. Increasing the concrete strength can be a cost-effective means of increasing shear strength. For exam- ple, increasing fc' from 27.6 MPa to 34.5 MPa (4000 psi to 5000 psi) yields a 12% increase in shearing resistance. In- creasing the diameter of a rock socket can add considerably to the cost, depending on rock type, drillability, socket depth, etc. Rock of higher strength, which is likely to coincide with the case when shear is critical, can be some of the most ex- pensive rock to drill. However, increasing the diameter can provide other benefits that may offset additional costs, such as reducing the congestion of reinforcement steel (improved constructability), increasing axial and bending capacity, and further limiting displacements. Shear strength of the shaft can also be increased by providing additional transverse reinforcement in the form of either spiral or ties. From Eq. 145, this can be achieved by in- creasing the size of transverse reinforcement or by decreas- ing the pitch(s). Constructability can be affected when bar spacings are too small to allow adequate flow of concrete. One aspect of reinforced-concrete behavior in shear that is not taken into account in any building code is confining d B Bls≈ +⎡⎣⎢ ⎤ ⎦⎥0 9 2. π

stress. Shear capacity of concrete is increased at higher confining stress and deep foundations are subjected to sig- nificant confinement, especially when they are embedded in rock. This is a topic that warrants research but has yet to be investigated in a meaningful way that can be applied to foundation design. Axial When lateral loading is not significant, structural design of concrete shafts must account for axial compression or ten- sion (e.g., uplift) capacity. For shafts designed for signifi- cant load transfer at the base, compression capacity of the reinforced-concrete shaft could be less than that of the rock bearing capacity. In high-strength intact rock, compressive strength of the shaft may be the limiting factor. For design of reinforced-concrete columns for axial compression, the AASHTO-factored axial resistance is given by (147) in which Pr = factored axial resistance, with or without flex- ure; φ = resistance factor (0.75 for columns with spiral trans- verse reinforcement, 0.70 for tied transverse reinforcement); fc' = strength of concrete at 28 days; Ag = gross area of the section; Ast = total area of longitudinal reinforcement; and fy = specified yield strength of reinforcement. One source of uncertainty is that the design equations given here are for unconfined reinforced-concrete columns. The effect of con- finement provided by rock on the concrete strength is not easy to quantify, but increases the strength compared with zero confinement, and warrants further investigation. SUMMARY Lateral loading is a major design consideration for trans- portation structures and in many cases governs the design of rock-socketed drilled shafts. Design for lateral loading must satisfy performance criteria with respect to (1) structural re- sistance of the reinforced-concrete shaft for the strength limit state and (2) deflection criteria for the service limit state. Analytical methods that provide structural analysis of deep foundations while accounting for soil–structure interaction have, therefore, found wide application in the transportation P f A A f Ar c g st y st= −( ) +[ ]φ0 85 0 85. . ' 70 field. However, the ability of analytical methods to account properly for rock mass response and rock–structure interac- tion has not developed to the same level as methods used for deep foundations in soil. The survey shows that most state DOTs use the program COM624 or its commercial version LPILE for design of rock-socketed shafts. Review of the p-y curve criteria cur- rently built into these programs for modeling rock mass re- sponse shows that they should be considered as “interim” and that research is needed to develop improved criteria. Some of this work is underway and research by North Carolina (Gabr et al. 2002), Ohio (Liang and Yang 2006), Florida (McVay and Niraula 2004), and Ashour et al. (2001) is described. All of these criteria are in various stages of development and are not being applied extensively. Models based on elastic continuum theory and developed specifically for rock-socketed shafts have been published. Two methods reviewed in this chapter are the models of Carter and Kulhawy (1992) and Zhang et al. (2000). Advan- tages and disadvantages of each are discussed and compared with p-y methods of analysis. These models are most useful as first-order approximations of shaft lateral displacements for cases where the subsurface profile can be approximated as consisting of one or two homogeneous layers. For exam- ple, when a preliminary analysis is needed to develop trial designs that will satisfy service limit state deflection criteria, the method of Carter and Kulhawy can provide convenient solutions that can be executed by means of spreadsheet analysis. A disadvantage of these methods is that they do not directly provide solutions to maximum shear and moment, parameters needed for structural design, and they do not in- corporate directly the nonlinear properties of the reinforced- concrete shaft. Structural issues associated with rock-socketed shafts are reviewed. The concept of depth of fixity is shown to be a use- ful analytical tool providing a link between geotechnical and structural analysis of drilled shafts. A method for establish- ing depth of fixity is presented and its use in the design process is described. Other issues identified by the survey, including high shear in short sockets subjected to high moment loading and its implications for reinforced-concrete design, are addressed.

Next: Chapter Five - Construction and Field Testing »
Rock-Socketed Shafts for Highway Structure Foundations Get This Book
×
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

TRB’s National Cooperative Highway Research Program (NCHRP) Synthesis 360: Rock-Socketed Shafts for Highway Structure Foundations explores current practices pertaining to each step of the design process, along with the limitations; identifies emerging and promising technologies; examines the principal challenges in advancing the state of the practice; and investigates future developments and potential improvements in the use and design of rock-socketed shafts.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  6. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  7. ×

    View our suggested citation for this chapter.

    « Back Next »
  8. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!