Free access
TECHNICAL PAPERS
Feb 20, 2009

Importance of Lower-Bound Capacities in the Design of Deep Foundations

Publication: Journal of Geotechnical and Geoenvironmental Engineering
Volume 135, Issue 7

Abstract

There is generally a physical limit to the smallest possible capacity for a deep foundation. However, a lower bound on the capacity has rarely been accounted for in performing reliability analyses and developing reliability-based design codes. The objectives of this paper are to investigate the effect of having a lower-bound capacity on the reliability of a geotechnical engineering system and to propose a load and resistance factor design (LRFD) checking format that includes information on the lower-bound capacity in design. It is concluded that a lower-bound capacity can cause a significant increase in the calculated reliability for a geotechnical design even if it is an uncertain estimate. Two alternative LRFD formats that incorporate lower-bound capacities and that would not require substantive changes to existing codes are proposed. Real-world examples dealing with the design of onshore and offshore foundations indicate that the incorporation of a lower-bound capacity into design is expected to provide a more realistic quantification of reliability for decision-making purposes and therefore a more rational and efficient basis for design.

Introduction

The ability to manage risk in geotechnical engineering relies on a realistic assessment of the probability of failure for designs. A realistic assessment compares the absolute risk against what is considered to be acceptable, and is thus important in risk management. Due to the implementation of a load and resistance factor design (LRFD) approach in American Association of State Highway and Transportation Officials (AASHTO) design specifications (AASHTO 1994, 2004), there has been extensive research recently to develop resistance factors for foundations by performing reliability analyses with pile load test databases. Examples include the work by Barker et al. (1991), Withiam et al. (1997), Goble (1999), Liang and Nawari (2000), McVay et al. (2000, 2002, 2003), Zhang et al. (2001), Kuo et al. (2002), Kulhawy and Phoon (2002), Phoon et al. (2003), Paikowsky (2003), and Withiam (2003). Similar work was conducted in the late 1980s to support a LRFD code for offshore pile design (API 1993). Examples include Tang (1988, 1990), Hamilton and Murff (1992), Tang and Gilbert (1993), Horsnell and Toolan (1996), and Bea et al. (1999).
There is a general belief that the calculated probabilities of failure from conventional reliability analyses are not realistic and are primarily useful as nominal or relative values for comparison purposes (Simpson et al. 1981; Wu et al. 1989; Rodriguez et al. 1998). For example, in the last 3040years , observations of failure rates for offshore piles indicated that the frequencies of foundation failures are significantly lower than the calculated probabilities of failure (Horsnell and Toolan 1996; Aggarwal et al. 1996; Bea et al. 1999). Discrepancies between calculated probabilities of failure and actual failure rates of offshore piles have been mainly attributed to a conservative bias in the models used to predict the capacity of offshore piles.
Another possible reason that the calculated probabilities of failure are not realistic is related to how the tails of the probability distributions for the load and the capacity are modeled (Fig. 1). Most analyses use the mean and variance of an assumed, mathematically convenient probability distribution (for example, normal or lognormal) to model the tails. For example, in most of the reliability analyses for deep foundations, the foundation capacity is modeled using a lognormal distribution and the coefficient of variation (COV) for the capacity is relatively large, ranging from 0.3 to 1.0 (Tang 1988, 1990; Hamilton and Murff 1992; API 1993; Horsnell and Toolan 1996; Bea et al. 1999; McVay et al. 2000, 2002, 2003; Kulhawy and Phoon 2002; Phoon et al. 2003; AASHTO 2004). A lognormal distribution, with a lower tail that extends to zero, does not capture the realistic possibility that there is a physical lower bound for the capacity of geotechnical systems.
Fig. 1. Reliability analysis using conventional and bounded probability distributions for capacity
Najjar (2005) investigated the possibility of a lower-bound capacity for driven steel pipe piles in cohesive and cohesionless soils using two well-known databases with results of axial load tests (Olson and Dennis 1982). Physical models for predicting lower-bound pile capacities were proposed and used to estimate lower-bound capacities for 45 piles in clays and 36 piles in sands. The data analyzed indicated strong evidence of the existence of a lower-bound capacity and a need to incorporate lower-bound capacities in more realistic probability distributions for capacity. The ratio of the lower-bound capacity to the predicted capacity has an average of 0.55 and 0.50 for piles driven in cohesive and cohesionless soils respectively (Najjar 2005).
In geotechnical engineering applications, the left-hand tail of the capacity distribution governs the probability of failure since the uncertainty in the capacity is generally larger than the uncertainty in the load. As a result, the reliability is expected to be strongly affected by the existence of a lower-bound capacity even when the lower-bound capacity is significantly smaller than the predicted capacity. The objectives of this paper are to: (1) investigate the effect of having a lower-bound capacity on the reliability of a geotechnical engineering system; (2) propose a LRFD design-checking format that includes information on the lower-bound capacity in addition to the conventional design information; and (3) present real-world examples to illustrate how information about the lower-bound capacity can be used to optimize the design for foundations.

Effect of Lower-Bound Capacity on Reliability

Mathematical Formulation for Reliability

A convenient mathematical model to account for a lower-bound capacity is a mixed lognormal distribution (Fig. 2). For capacities greater than the lower bound, the distribution is a continuous probability density function that follows a lognormal distribution. Most reliability analyses for pile capacities have assumed lognormal distributions for the pile capacity based on the available database information (AASHTO 2004; McVay et al. 2000, 2002, 2003), and the model on Fig. 2 is consistent with this conventional approach. For capacities at the lower bound, there is a finite probability (that is, a probability mass function) that corresponds to the probability of being less than or equal to the lower bound in the nontruncated lognormal distribution. For driven piles in cohesive and cohesionless soils, mixed lognormal probability distributions provide a realistic model to the uncertainty in the capacity of piles, as observed in results of pile load tests (Najjar 2005).
Fig. 2. Mixed distribution for modeling capacity
When both the load and the capacity follow lognormal probability distributions, the probability of failure can be calculated analytically as
pf=Φ(ln(FSmedian)ln[(1+δS2)(1+δR2)])=Φ(β)
(1)
where Φ()=standard normal cumulative distribution function; FSmedian=median factor of safety defined as the ratio of the median capacity rmedian , to the median load smedian ; δR and δS=coefficients of variation of the capacity and load, respectively; and β=reliability index.
For the case where a mixed lognormal distribution is used to model the capacity and assuming a lognormal distribution for the load, the reliability and the probability of failure are calculated based on the parameters presented in Fig. 3 as
Reliability=1pf=P(S<R)=P(S<Rr=LB)P(R=LB)+LBP(S<Rr)fR(r)dr=Φ(ln(LB)ln(smedian)ln(1+δS2))Φ(ln(LB)ln(rmedian)ln(1+δR2))+LBΦ(ln(r)ln(smedian)ln(1+δS2))12πln(1+δR2)re12((ln(r)ln(rmedian))(ln(1+δR2)))2dr
(2)
where LB=lower -bound capacity. Note that the median and COV for the lognormal distribution of capacity, rmedian and δR , respectively, correspond to the nontruncated distribution. A more general form of Eq. (2) is obtained by normalizing the load and the capacity by the median capacity, rmedian
Reliability=1pf=P(S<R)=P(S<Rr=LB)P(R=LB)+LBP(S<Rr)fR(r)dr=Φ(ln(LBrmedian)ln(1FSmedian)ln(1+δS2))Φ(ln(LBrmedian)ln(1)ln(1+δR2))+LBrmedianΦ(ln(rrmedian)ln(1FSmedian)ln(1+δS2))e12((ln(rrmedian)ln(1))(ln(1+δR2)))22πln(1+δR2)rrmediand(rrmedian)
(3)
Hence, for given values of LBrmedian , δR , and δS , the reliability can be expressed entirely as a function of the median factor of safety. Note that the factor of safety used in design is not necessarily equal to the median factor of safety. If conservative (versus “best guess”) values of the load and capacity are used in calculating the design factor of safety, then the median factor of safety will be greater than the design factor of safety. When LBrmedian>0 , an analytical solution for Eq. (3) is not available and it must be solved using numerical integration.
Fig. 3. Factors that affect reliability of pile foundation

Reliability versus Lower-Bound Capacity

The probability of failure and the reliability of a pile foundation are expected to be affected by the lower-bound capacity, the median factor of safety, and the relative uncertainty in the load and capacity (Fig. 3). Curves showing the variation of the reliability index as a function of the ratio of the lower-bound to median capacity are shown on Fig. 4. The curves in Fig. 4 represent the case where the uncertainty in the capacity (δR=0.4) is relatively large compared to the uncertainty in the load (δS=0.2) . This example is representative of many geotechnical designs where the capacity is more uncertain than the load (McVay et al. 2000; Kulhawy and Phoon 2002; Phoon et al. 2003; AASHTO 2004).
Fig. 4. Effect of lower-bound capacity on reliability ( δR=0.4 and δQ=0.2 )
The primary conclusion from Fig. 4 is that a lower-bound capacity can have a significant effect on the calculated reliability. For example, consider a typical case where the median factor of safety is 3.0. If the lower-bound capacity is anything greater than 0.6 times the median capacity, the probability of failure is reduced by more than an order of magnitude compared to the case where there is no lower bound (Fig. 4). The effect of the lower bound capacity on the reliability is influenced by the magnitude of the median factor of safety; as the median factor of safety increases, the lower bound is more significant in reducing the probability of failure.
To better illustrate the magnitude of the effect of the lower-bound capacity, the median factor of safety that is required to achieve different levels of reliability is plotted on Fig. 5 as a function of the ratio of the lower-bound to the median capacity. The required median factor of safety decreases as the lower-bound capacity approaches the median capacity. Results in Fig. 5 indicate that resistance factors in a LRFD should incorporate information about the lower-bound capacity if they are to provide a consistent level of reliability. The reliability index of a design that is based on conventional resistance factors may differ significantly from the reliability index obtained when the effect of the lower-bound capacity is accounted for in the design-checking equations.
Fig. 5. Equivalent factors of safety for different lower-bound values ( δS=0.2 , δR=0.4 )

Effect of Probability Distribution for Capacity

In this section, the effect of varying the distributional form for capacity on the calculated reliability is investigated by considering a variety of bounded probability distributions. These include truncated normal and lognormal, uniform, and beta distributions, in addition to the mixed lognormal distribution used to generate Figs. 4 and 5. Mathematical expressions for the probability of failure and the probability density function for each distribution are presented in the Appendix. In all cases, the uncertainty in the load is modeled using a conventional lognormal distribution. The variation of the reliability index with the ratio of the lower-bound to median capacity for different capacity distributions is plotted in Fig. 6.
Fig. 6. Reliability calculations for different capacity distributions ( δS=0.2 , δR=0.4 )
Results in Fig. 6 indicate that for a given ratio of lower-bound to median capacity, the probability distribution for capacity does not have a significant effect on the calculated reliability, especially as the median factor of safety increases. This result is important because it shows that the reliability of a design is significantly more affected by the ratio of the lower-bound to median capacity than it is by the type of probability distribution used to model the capacity. It is worth noting that reliability indices that are calculated using the mixed lognormal distribution tend to be smaller than those obtained using other capacity distributions, indicating that the effect of the lower-bound capacity on the reliability can be higher than that shown in Figs. 4 and 5.

Simplified Bilinear Reliability Model

Curves in Fig. 4 indicate that there is a threshold value for the ratio of the lower-bound to the median capacity above which the lower bound affects the reliability (the knee in the curves in Fig. 4). Below this threshold, the lower-bound capacity has essentially no effect on the reliability index. Above this threshold, the reliability index increases approximately linearly with increasing lower bound. The threshold decreases as the median factor of safety increases. A proposed approximation for this threshold is the “most probable failure point” for the capacity in the case where there is no lower bound, which is given by Ang and Tang (1984) for the case of a lognormal capacity and load distributions as
LBthresholdrmedianrMPP,LB=0rmedian=e([ln(1+δR2)]ln[(1+δS2)(1+δR2)])β
(4)
Curves in Fig. 4 also indicate that the shape of the curves for relatively large ratios of lower-bound to median capacities can be approximated analytically by assuming that the capacity is equal to the lower-bound value. For a load following a lognormal distribution, the probability of failure is bounded by the following analytical expression:
pfP(S>LB)=Φ(ln(FSmedian)+ln(LBrmedian)ln(1+δS2))
(5)
This bounding approximation is shown on Fig. 4. As the ratio of the lower-bound to median capacity approaches one, the curves approach the bounding approximation in Eq. (5). A noteworthy aspect of this approximation is that it does not depend on an assumed probability distribution for capacity.
Based on the observations presented above, a simple approximation is proposed to relate the reliability to the lower-bound capacity (Fig. 7). The model is defined by three points that can be evaluated without the need for numerical integration. First, the reliability index with no lower-bound capacity is calculated from Eq. (1). Then a horizontal line is drawn from zero to the most probable failure point defined by Eq. (4). Finally, the reliability index with LBrmedian=1.0 is calculated from Eq. (5) and a straight line is drawn from the reliability index with no lower-bound capacity at the threshold to this reliability index at LBrmedian=1.0 .
Fig. 7. Proposed bilinear model for evaluating reliability
Bilinear approximations to the curves describing the variation of the reliability index with the lower-bound capacity are shown in Fig. 8. Reliability curves corresponding to a mixed lognormal capacity distribution are also shown in Fig. 8. The bilinear model provides a conservative but close approximation to the calculated reliability without the need for complicated computations.
Fig. 8. Reliability approximations with proposed bilinear model ( δS=0.2 , δR=0.4 )

Effect of Uncertainty in Lower-Bound Capacity on Reliability

In practice, the lower-bound capacity may not be known with certainty. For example, spatial variability in soil properties or inaccuracies in a proof-load test can lead to uncertainty in estimating a lower-bound capacity. In this section, a mathematical framework is presented for incorporating information about uncertainty in lower-bound capacities in reliability analyses.
If uncertainty in the lower-bound capacity exists, the probability of failure is calculated using the theorem of total probability as
Pf=(pfLB)fLB(LB)d(LB)
(6)
where pfLB=probability of failure given a lower-bound capacity LB; and fLB(LB)=probability density function of the lower-bound capacity (LB). To illustrate the effect of uncertainty in the lower-bound capacity on the reliability, the following case is considered: the load (S) follows a lognormal distribution with a COV of 0.2 and the capacity (R) follows a mixed lognormal distribution with a nontruncated COV of 0.4 and conditional on a lower-bound capacity (LB) that follows a lognormal distribution with a range of COV values.
The effect of uncertainty in the lower-bound capacity on the reliability index is shown in Fig. 9. Even when there is uncertainty in the lower-bound capacity, the presence of a lower-bound capacity increases the reliability compared to the conventional case where a lower-bound capacity is not included in the reliability analysis. Furthermore, the larger the ratio of the median lower-bound to the median capacity is, the greater the increase in reliability. When the lower bound is relatively close to the median capacity (i.e., the ratio of median lower bound to median capacity equal to 0.5 and 0.6 in Fig. 9), increasing uncertainty in the lower-bound value decreases the reliability. This result occurs because the possibility of smaller values for the lower bound tends to decrease the reliability. However, when the lower bound is relatively far from the median capacity (i.e., the ratio of the median lower-bound to median capacity equal to 0.4 in Fig. 9), increasing uncertainty in the lower bound actually causes a small increase in the reliability. This result occurs when the possibility becomes dominant where the lower bound is less than the threshold where the reliability is affected by it (i.e., to the left of the knees in the curves in Fig. 4).
Fig. 9. Effect of uncertainty in lower-bound capacity on reliability ( FSmedian=3.0 , δS=0.2 , δR=0.4 )
For practical applications, uncertainties in lower-bound capacities are expected to be small relative to uncertainties about mean capacities. For example, a lower-bound capacity that is calculated based on lower-bound shear strength values and soil properties (Najjar 2005) is expected to be less sensitive to variability in the depositional environment, in situ soil structure, soil-structure interaction, and sampling and testing techniques. As a result, lower-bound capacities may be estimated with a higher degree of confidence compared to expected or mean capacities.
To illustrate and test this hypothesis, the effect of spatial variability in the undrained shear strength on the predicted (or mean) and lower-bound pile capacities is investigated for a number of offshore sites. Predicted capacities of piles in normally consolidated to slightly overconsolidated clay are generally obtained using undisturbed undrained shear strength properties while estimates of lower-bound pile capacities can be obtained using remolded undrained shear strength values using the model presented in Najjar (2005). Both vertical and horizontal variations in shear strength properties will affect the predicted and lower-bound capacities of piles.
To study the effect of vertical variations on the undisturbed and remolded shear strength, shear strength measurements obtained at two offshore sites in the Gulf of Mexico (Doyle 1999) are presented in Fig. 10. The data were obtained from unconsolidated, undrained triaxial shear tests on undisturbed and remolded samples. The following two observations can be made by comparing the shear strength profiles at the two sites: (1) the variability of the undisturbed shear strength with depth is much larger than that of the remolded shear strength, indicating that uncertainty in predicted pile capacity due to vertical variations in soil properties is expected to be larger than that of the lower-bound capacity; and (2) the remolded shear strength profile provides a lower bound to variations about the undisturbed profile.
Fig. 10. Variability in remolded and undisturbed undrained shear strength profiles at two borings
To study the effect of horizontal variations on undisturbed and remolded shear strengths, strength measurements obtained from four different boreholes separated by about 6km in an offshore site (Al-Awar 2002) are plotted in Fig. 11. The shear strength was measured using unconsolidated, undrained triaxial shear tests on undisturbed and remolded specimens. The variability in the measured strengths reflects both vertical and horizontal variations. In order to isolate the contribution of horizontal variations, the strength data were averaged vertically to different depths in the four borings. This vertically averaged strength is of interest in pile design since the side friction will be approximately proportional to it. The horizontal variations in the vertically averaged shear strength over varying depths are compared in Fig. 11 and in Table 1. The results indicate that horizontal variations in the remolded shear strength are significantly lower than those in the undisturbed strength and can be represented by a coefficient of variation of about 0.1–0.15. This small level of uncertainty in the lower-bound capacity will not appreciably reduce the impact of lower-bound capacities on the reliability, as indicated by Fig. 9.
Fig. 11. Variability in remolded and undisturbed undrained shear strength from four borings
Table 1. Comparison of Horizontal Variations in Remolded and Undisturbed Shear Strength
Depth(m)Vertically averagedundisturbed shear strengthVertically averaged remolded shear strength
Standarddeviation(kPa)Coefficient of variationStandarddeviation(kPa)Coefficient of variation
156.60.272.10.27
309.50.261.90.15
4512.50.261.20.08
6016.00.262.00.11

Incorporating Lower-Bound Capacities into LRFD

Since a lower-bound capacity can have a significant effect on the reliability of a design, a reliability-based LRFD design code could include information on the lower-bound capacity. Two alternative formats are proposed here for including information about a lower-bound capacity in a LRFD design code: (1) a conventional design checking equation where the resistance factor is adjusted according to the lower-bound capacity; and (2) a second design checking equation to include information about the lower-bound capacity. Both design formats require that a lower-bound capacity be properly determined using physical models (Najjar 2005) that are based on site-specific soil properties. Since for practical applications, uncertainties in lower-bound capacities are expected to be relatively small, the proposed LRFD design-checking formats will not include the effect of uncertainties in lower-bound capacities in the analysis.
It should be noted that the proposed LRFD design-checking formats which include information about the lower-bound capacity correspond to the ultimate limit state only. Checking of settlement or serviceability requirements when a lower-bound capacity is included in the design is required but is beyond the scope of this paper.

Adjusted Resistance Factor for Lower-Bound Capacity

The conventional design checking equation has the following general form:
φRrnominalγSsnominal
(7)
where rnominal=nominal capacity calculated using a design method; φR=resistance factor; snominal=nominal load for design; and γS=load factor. In order to incorporate the effect of a lower-bound capacity, this design checking equation is modified as follows:
φR(LB)rnominalγSsnominal
(8)
where the resistance factor, φR(LB)=function of the lower-bound capacity. When there is no lower bound and when the load and the capacity follow conventional lognormal distributions, the reliability is calculated from Eq. (9) as
Reliability=Φ(ln(FSmedian)ln[(1+δS2)(1+δR2)])ifLB=0
(9)
where FSmedian=median factor of safety; and δS and δR=coefficients of variation of the load and the capacity, respectively. Assuming λR and λS to be mean biases for the capacity and load respectively (mean value divided by the nominal value), the nominal values of the load and the capacity will thus be related to the median load and median capacity through
rmedian=λRrnominal1+δR2andsmedian=λSsnominal1+δS2
(10)
When there is no lower bound and the reliability is found from Eq. (9), the required value for the resistance factor can be expressed in terms of the target reliability index as follows:
φR=γSeβtargetln[(1+δS2)(1+δR2)](λRλS)1+δS21+δR2ifLB=0
(11)
If the nominal values of the load and the capacity are assumed to be equal to the mean values (the methods for predicting the load and capacity are unbiased), the required value for the resistance factor can be expressed in terms of the target reliability index as follows:
φR=γSeβtargetln[(1+δS2)(1+δR2)]1+δS21+δR2ifLB=0
(12)
The required resistance factor can also be expressed in terms of the required median factor of safety as
φR=γSFSmedian(λRλS)1+δS21+δR2ifLB=0
(13)
If the methods for predicting the load and capacity are unbiased
φR=γSFSmedian1+δS21+δR2ifLB=0
(14)
For a nonzero lower-bound capacity, the required median factor of safety will decrease as the lower bound increases. Since the lower-bound capacity does not affect the other variables in Eqs. (13) - (14), the required resistance factor when there is a lower-bound capacity, φR(LB) , can be expressed in terms of the conventional case as follows:
φR(LB)=(FSmedian(LB=0)FSmedian(LB))φR
(15)
where φR is obtained from Eqs. (13) - (14) and corresponds to the case where there is no lower-bound capacity; FSmedian(LB=0)=median factor of safety required for a target reliability index if there is no lower bound [obtained from Eq. (9)]; and FSmedian(LB)=median factor of safety required to achieve the target reliability index for a lower-bound value equal to LB. Therefore, the increase in the resistance factor due to an increase in the lower-bound capacity is inversely proportional to the corresponding decrease in the required value for the median factor of safety.
The variation of the ratio of the resistance factor incorporating a lower-bound capacity to the conventional resistance factor, φR(LB)φR , is shown as a function of the lower-bound capacity in Fig. 12 for different target values of the reliability index. For reasonable values of the ratio of the lower-bound to median capacity, 0.4–0.9, the effect of the lower bound on the required resistance factor is significant. Consider a design with a target reliability index of 3.0 and a COV value for the capacity of 0.5. If the lower-bound capacity is 0.7 times the median capacity, then the required resistance factor in the design-checking equation, Eq. (8), will be more than double the conventional resistance factor [Fig. 12(b)]. For example, a conventional resistance factor of 0.25 would be increased to 0.6 in accounting for the lower-bound capacity.
Fig. 12. Variation of increase in nominal resistance factor with lower-bound capacity (COVLoad=δS=0.15)

Added Design Checking Equation for Lower-Bound Capacity

An alternative code format would be to have two design checking equations
φRrnominalγSsnominal
φLBLBγSsnominal
where the first design-checking equation is the conventional equation and the second equation includes a resistance factor, φLB , that is applied directly to the lower-bound capacity. A design will be considered to be acceptable (specified level of reliability is achieved) if only one of the two design equations is satisfied. The motivation for this form of the design-checking equation is that the conventional approach is incorporated and does not need to be modified, whether or not there is a lower-bound capacity; the effect of a lower-bound capacity is reflected entirely in the second equation.
The required value for φLB can be determined such that the target reliability is achieved by rearranging Eq. (13) so that it is expressed in terms of the lower-bound capacity rather than the nominal capacity
φLB=γS(LBrnominal)FSmedian(LB)(λRλS)1+δS21+δR2
(17)
The nominal capacity, rnominal , is related to the median capacity, rmedian , through Eq. (10). As such, the lower-bound resistance factor given by Eq. (17) can be expressed in terms of the ratio of the lower-bound to median capacity as
φLB=γSλS(LBrmedian)FSmedian(LB)1+δS2
(18)
If the methods for predicting the load and capacity are unbiased
φLB=γS(LBrmedian)FSmedian(LB)1+δS2
(19)
A plot of φLB versus the lower-bound capacity is shown in Fig. 13 for different target reliability indices. Approximations of φLB as calculated using the simple bilinear reliability model that was proposed earlier are also plotted in Fig. 13 for comparison. The curves begin at values of the lower-bound capacity where the second design checking equation in Eq. (16) governs. One advantage of this approach with two design-checking equations [Eq. (16) versus Eq. (8)] is that φLB is not very sensitive to either the magnitude of the lower-bound capacity or the target reliability index (Fig. 13). In fact, a conservative value of around 0.75 for φLB could be used to cover a wide range of possibilities.
Fig. 13. Variation of lower-bound resistance factor to account for lower-bound capacity (COVLoad=δS=0.15)

Illustrative Example 1: Design of Bridge Foundation

In this section, a case study presented by Goble (1996) is used to illustrate how the proposed LRFD checking equation that incorporates a lower-bound capacity can be used in the design of a bridge foundation. The foundation to be designed supports an isolated bridge column that is subjected to a critical factored axial load of 10MN . The piles to be used in supporting the bridge column consist of closed-ended steel pipe piles with an outside diameter of 355mm and a length of 25m . A simplified schematic of the soil profile at the site is shown in Fig. 14.
Fig. 14. Soil profile for case study example (from Goble 1996)
The predicted nominal axial capacity of a single driven pile is 1.8MN as calculated by the API method (API 1993). If the COV values for the capacity (δR) and load (δS) are 0.5 and 0.15, respectively, which are typical for driven piles in cohesionless soils and for bridge loading conditions, respectively, the mean biases ( λR and λS ) for the capacity and load are 1.0, and the target reliability index is 3.0, then the conventional resistance factor (φR) is obtained from Eq. (12) and is equal to 0.23. The factored capacity per pile is calculated from Eq. (8): φRrnominal=0.23×1.8MN=0.42MN . Since each pile provides a factored capacity of 0.42MN and the total factored load is 10MN , the total number of piles needed to support the column load is 10MN0.42MN=24 piles based on the conventional approach assuming a group efficiency factor of 1.0.
To account for the effect of a lower-bound capacity in this design, a lower-bound capacity is calculated for each pile using the available soil profile (Fig. 14) and a lower-bound model that is consistent with driven piles in siliceous, cohesionless soils as presented by Najjar (2005). In this model, the lower-bound skin friction is calculated assuming that the lateral coefficient of earth pressure is replaced with the at-rest value and the soil-pile friction angle and end-bearing capacity factors are replaced with the values for one category less in density (e.g., the values for a “dense sand” are replaced with those for a “medium sand”). The skin friction is thus calculated as kLBσvtanδLB=0.55×145kNm2×tan(20)=29kNm2 , where kLB is equal to 0.55 (at rest value), and δLB is equal to 20° (API value for loose sand and gravel assuming one category less in density). The skin friction in the soft sandy silty clay layer (SPT=6) in the top 6m is neglected. The lower-bound capacity of the pile due to skin friction is thus equal to 0.61MN . The lower-bound tip resistance can also be calculated assuming one category less in density as σvNq,LB=242.7kNm2×12=2,870kNm2 , resulting in a lower-bound end bearing capacity of 0.29MN .
The calculated lower-bound capacity is thus equal to 0.9MN , or 0.5 times the predicted capacity and 0.55 times the median capacity. From Fig. 12(b), the resistance factor accounting for the lower bound can be increased by 1.7 times the conventional value of 0.23: φR(LB=0.5rnominal)=1.7×0.23=0.40 . The corresponding factored capacity per pile is then found from Eq. (8): φR(LB=0.5rnominal)rnominal=0.40×1.8MN=0.72MN .
Alternatively, the lower-bound resistance factor can be found from Fig. 13(b) to be φLB=0.81 . According to Eq. (16), the target reliability will be achieved with the larger of the factored nominal capacity from the conventional equation, φRrnominal=0.42MN , and the factored capacity based on the lower-bound capacity, φLBLB=0.81×0.9MN=0.72MN . Therefore, the same factored capacity of 0.72MN for design is obtained using either version of the design checking equation [Eq. (8) or (16)], and the required number of piles to support the column load is 10MN0.72MN=14 . Incorporating the lower-bound capacity into this design has a significant effect on the level of conservatism required, reducing the required number of piles from 24 to 14 while still providing the target level of reliability for each foundation.

Illustrative Example 2: Design of Offshore Foundation

In this section, an illustrative example is provided whereby a lower-bound capacity is incorporated into a reliability-based design. The example is based on a case history for the design of a floating offshore facility for oil production. The target annual reliability index for the design of the foundations is set to 3.0. Due to the geometry of the structure and the metocean environment, the uncertainty in the load is relatively low (COV=0.2) . Conversely, the uncertainty in the capacity is relatively large (COV=0.4) because the site is located in a frontier area with very little available data.
From a conventional reliability analysis where the load and the capacity are assumed to follow lognormal distributions, a median factor of safety greater than 3.5 is needed to achieve the target reliability index of 3.0 (Fig. 15). For this example design, the median load is 10MN and the required median capacity is 37MN . The opportunity for optimization in this design is that full setup of the pile capacity is estimated to take about 1year . Therefore, oil production cannot begin until at least 1year after the piles are installed. If a lower-bound estimate of capacity based on pile installation data is incorporated into the reliability analysis, it may be possible to accelerate the time that oil production can begin.
Fig. 15. Effect of lower-bound capacity on foundation reliability
Fig. 15 shows how a lower-bound capacity affects the reliability index. In this case, the estimated capacity from a re-tap analysis conducted several days after pile driving was 18MN , or about 50% of the median capacity after full setup. The results in Fig. 15 show that a median factor of safety of 2.7 or a median capacity of 27MN , is sufficient to provide an adequate level of reliability. The setup analysis indicates that this capacity will be achieved within 3months after installation. Therefore, the oil-production schedule can be moved up 9months by conducting a more realistic reliability assessment that incorporates information about the lower-bound capacity.

Conclusions

The major conclusion from this work is that a lower-bound capacity can cause a significant increase in the calculated reliability for a geotechnical design even when the lower-bound capacity is significantly smaller than the mean or median capacity. The effect of the lower-bound capacity on the reliability is most pronounced when the uncertainty in the capacity is large compared to that in the load and when the degree of conservatism in the design is large. Results from reliability analysis also indicated that the reliability of a geotechnical system can be more sensitive to the lower-bound capacity than to the median capacity. This signifies that: (1) resistance factors in a LRFD should incorporate information about the lower-bound capacity if they are to provide a consistent level of reliability; and (2) reliability analyses will provide more realistic and useful information for decision-making purposes if they include information about lower-bound capacities.
Reliability assessments using different distributions for capacity indicated that the reliability of a design is affected more by the ratio of the lower-bound to median capacity than it is by the type of probability distribution used to model the capacity. Based on the observed variation of the probability of failure with the ratio of the lower-bound to median capacity, a simple bilinear model is proposed to relate the reliability to the lower-bound capacity as an approximation. Reliability assessments using the proposed model provided a close approximation to the reliability obtained using more complicated methods that are based on numerical integration. The proposed model provides an effective and simple tool for the practicing engineer to solve practical reliability-based problems in a realistic manner.
Two LRFD design-checking formats that include information on lower-bound capacities in addition to the conventional design information are proposed and methodologies for calculating resistance factors are presented. The effect of the lower-bound capacity on design for both formats is reflected in the magnitudes of the resistance factors. The applicability of the two proposed LRFD approaches is illustrated using a real-world example dealing with the design of a bridge foundation. Incorporating the lower-bound capacity into this design had a significant effect on the level of conservatism required, reducing the required number of piles from 24 to 14 while still providing the target level of reliability for each foundation. A real-world example dealing with the design of an offshore foundation for an oil production facility indicated that if a lower-bound estimate of capacity based on a pile re-tap analysis is incorporated into the reliability analysis, it may be possible to accelerate the time that oil production can begin by 9months , while maintaining a design with an adequate level of reliability.

Notation

The following symbols are used in this paper:
FSmedian
=
median factor of safety;
FX(x)
=
cumulative probability distribution function of X ;
fX(x)
=
probability distribution function of X ;
LB
=
lower-bound capacity;
pf
=
probability of failure;
pX(x)
=
probability mass function of X ;
r
=
capacity or resistance;
rmedian
=
median capacity;
s
=
load;
β
=
reliability index;
γS
=
load factor;
δR
=
coefficient of variation of capacity;
δs
=
coefficient of variation of load;
Φ()
=
standard normal cumulative distribution function;
λ  and  ζ
=
parameters of lognormal distribution; and
φR
=
resistance factor.

Appendix. Probability Distributions

In this Appendix, mathematical expressions for the probability density functions for capacity are presented together with expressions for the probability of failure. The probability distributions include truncated lognormal, truncated normal, uniform, and beta distributions.

Truncated Normal and Lognormal Distributions

Mathematical expressions for the probability density functions and the probability of failure for truncated normal and lognormal distributions are presented in Eqs. (20) - (23), respectively
fR(r)=[11Φ(LBμRσR)]12πσRexp12(rμRσR)2
(20)
fR(r)=[11Φ(ln(LB)λRξR)]12πξRrexp(12)(ln(r)λRξR)2
(21)
pf=LB+[12πsξSe(12)(lnsλSξS)2][11Φ(LBμRσR)LBs12πσRe(12)(rμRσR)2dr]ds
(22)
pf=LB+[12πsξSe(12)(lnsλSξS)2][11Φ(lnLBλRξR)LBs12πrξRe(12)(lnrλRξR)2dr]ds
(23)

Uniform Distribution

The uniform distribution is defined by a probability density function that has a constant value over a range defined by a lower-bound capacity LB, and an upper-bound capacity UB, and is expressed as
fR(r)=1UBLBforLBrUB
fR(r)=0elsewhere
The expression for the probability of failure for the case where the capacity follows a uniform distribution is given by Eq. (25)
pf=LB+[12πsξSe12(lnsλSξS)2][LBs1UBLBdr]ds
(25)

Beta Distribution

A beta distribution with a mean μR and a standard deviation σR is defined over a range LB to UB and has the following probability density function:
fR(r)=1UBLB(Γ(α+β)Γ(α)Γ(β))(rLBUBLB)(α1)(UBrUBLB)(β1)
(26)
where
α=X2Y2(1X)X,β=α(1X)X,X=μRLBUBLB,Y=σRUBLB
and Γ()=gamma function defined as Γ(k)=0xk1exdx for k>0 . Depending on the values and signs of the parameters α and β , the beta distribution can assume different geometrical shapes.
The expression for the probability of failure for the case where the capacity follows a beta distribution is given by Eq. (27)
pf=LB+[12πsξSe(12)(lnsλSξS)2][LBs1UBLBΓ(α+β)Γ(α)Γ(β)(rLBUBLB)(α1)(UBrUBLB)(β1)dr]ds
(27)

References

AASHTO. (1994). LRFD bridge and construction specifications, Washington, D.C.
AASHTO. (2004). LRFD bridge design specifications, Washington, D.C.
Aggarwal, R. K., Litton, R. W., Cornell, C. A., Tang, W. H., Chen, J. H., and Murff, J. D. (1996). “Development of pile foundation bias factors using observed behavior of platforms during hurricane Andrew.” Proc., Offshore Technology Conf., OTC 8078, Houston, 445–455.
Al-Awar, Z. (2002). “Developing a model for axial pile capacity in an offshore field.” Master’s thesis, Univ. of Texas at Austin, Austin, Tex.
American Petroleum Institute (API). (1993). API RP 2A-LRFD: Recommended practice for planning, designing and construction fixed offshore platforms—Load and resistance factor design, 1st Ed., Washington, D.C.
Ang, A. H.-S., and Tang, W. H. (1984). Probability concepts in engineering planning and design. Vol. II. Decisions, risk and reliability, Wiley, New York.
Barker, R. M., Duncan, J. M., Rojiani, K. B., Ooi, P. S. K., Tan, C. K., and Kim, S. G. (1991). “Load factor design criteria for highway structure foundations.” Final Rep. No. NCHRP Proj. 24-4, Virginia Polytechnic Institute and State Univ., Blacksburg, Va.
Bea, R. G., Jin, Z., Valle, C., and Ramos, R. (1999). “Evaluation of reliability of platform pile foundations.” J. Geotech. Geoenviron. Eng., 125(8), 696–704.
Doyle, E. H. (1999). “Pile installation performance for four TLP’s in the Gulf of Mexico.” Proc., 31st Offshore Technology Conf., OTC 10826, Houston.
Goble, G. G. (1996). “Load and resistance factor design of driven piles.” Transportation Research Record. 1546, Transportation Research Board, Washington, D.C., 88–93.
Goble, G. G. (1999). “Geotechnical related development and implementation of load and resistance factor design (LRFD) methods.” NCHRP synthesis of highway practice 275, TRB, Washington, D.C.
Hamilton, J. M., and Murff, J. D. (1992). “Selection of LRFD resistance factors for pile foundation design.” Proc., Structural Congress ’92, ASCE, San Antonio, 788–795.
Horsnell, M. R., and Toolan, F. E. (1996). “Risk of foundation failure of offshore jacket piles.” Proc., Offshore Technology Conf., OTC 7997, Houston, 381–392.
Kulhawy, F. H., and Phoon, K. K. (2002). “Observations on geotechnical reliability-based design development in North America.” Proc., Int. Workshop on Foundation Design Codes and Soil Investigation in View of International Harmonization and Performance Based Design, Hayama, Japan, 31–48.
Kuo, C. L., McVay, M. C., and Birgisson, B. (2002). “Calibration of load and resistance factor design.” Transportation Research Record. 1808, Transportation Research Board, Washington, D.C., 108–111.
Liang, R., and Nawari, N. O. (2000). “Evaluation of resistance factors for driven piles.” Geotechnical Special Publication No. 100, ASCE, New York, 178–191.
McVay, M. C., Birgisson, B., Nguygen, T., and Kuo, C. L. (2002). “Uncertainty in load and resistance factor design phi factors for driven prestressed concrete piles.” Transportation Research Record. 1808, Transportation Research Board, Washington, D.C., 99–107.
McVay, M. C., Birgisson, B., Zhang, L., Perez, A., and Putcha, S. (2000). “Load and resistance factor design (LRFD) for driven piles using dynamic methods—A Florida perspective.” Geotech. Test. J., 23(1), 55–66.
McVay, M. C., Ellis, R. D., Birgisson, B., Consolazio, G. R., Putcha, S., and Lee, S. M. (2003). “Load and resistance factor design, cost, and risk.” Transportation Research Record. 1849, Transportation Research Board, Washington, D.C. 98–106.
Najjar, S. S. (2005). “The importance of lower-bound capacities in geotechnical reliability assessments.” Dissertation, Univ. of Texas at Austin, Austin, Tex.
Olson, R. E., and Dennis, N. D. (1982). “Review & compilation of pile test results, axial pile capacity.” PRAC Project 81-29, American Petroleum Institute, Dallas.
Paikowsky, S. G. (2003). “Practical lessons learned from applying the reliability methods to LRFD for the analysis of deep foundations.” Proc., Int. Workshop on Limit State Design in Geotechnical Engineering Practice (LSD 2003), Mass.
Phoon, K. K., Kulhawy, F. H., and Grigoriu, M. D. (2003). “Development of a reliability-based design framework for transmission line structure foundations.” J. Geotech. Geoenviron. Eng., 129(9), 798–806.
Rodriguez, S. G. H., Henriques, C. C. D., Ellwanger, G. B., and de Lima, E. C. P. (1998). “Structural reliability with truncated probability distributions.” Proc., Int. Conf. on Offshore Mechanics and Arctic Engineering, OMAE, Lisbon, Portugal.
Simpson, B., Pappin, J. W., and Croft, D. D. (1981). “An approach to limit state calculations in geotechnics.” Ground Eng., 14(6), 21–28.
Tang, W. H. (1988). “Offshore axial pile design reliability.” Research Rep. to API, Phase I, PRAC Project No. 86-29B, Univ. of Illinois, Urbana-Champaign, Ill.
Tang, W. H. (1990). “Performance reliability of offshore piles.” Proc., Offshore Technology Conf., OTC 6379, Houston, 299–308.
Tang, W. H., and Gilbert, R. B. (1993). “Case study of offshore pile system reliability.” Proc., Offshore Technology Conf., OTC 7196, Houston, 677–686.
Withiam, J. L. (2003). “Implementation of the AASHTO LRFD bridge design specifications for substructure design.” Proc., Int. Workshop on Limit State Design in Geotechnical Engineering Practice (LSD 2003), Boston.
Withiam, J. L., Voytko, E. P., Barker, R. M., Duncan, J. M., Kelly, B. C., Musser, S. C., and Elias, V. (1997). “Load and resistance factor design (LRFD) for highway bridge substructures.” Rep. No. DTFH61-94-C-00098, FHwA, U.S. Department of Transportation, Washington, D.C.
Wu, T. H., Tang, W. H., Sangrey, D. A., and Baecher, G. B. (1989). “Reliability of offshore foundations—State of the art.” J. Geotech. Engrg., 115(2), 157–178.
Zhang, L., Tang, W. H., and Ng, W. W. (2001). “Reliability of axially loaded driven pile groups.” J. Geotech. Geoenviron. Eng., 127(12), 1051–1060.

Information & Authors

Information

Published In

Go to Journal of Geotechnical and Geoenvironmental Engineering
Journal of Geotechnical and Geoenvironmental Engineering
Volume 135Issue 7July 2009
Pages: 890 - 900

History

Received: Apr 3, 2008
Accepted: Nov 11, 2008
Published online: Feb 20, 2009
Published in print: Jul 2009

Permissions

Request permissions for this article.

Authors

Affiliations

Shadi S. Najjar, A.M.ASCE [email protected]
Assistant Professor, Dept. of Civil and Environmental Engineering, American Univ. of Beirut. P.O. Box 11-0236, Riad El-Solh 1107-2020, Beirut (corresponding author). E-mail: [email protected]
Robert B. Gilbert, M.ASCE [email protected]
Professor, Civil Engineering Dept., Univ. of Texas at Austin, ECJ 9.227, Austin, TX 78712. E-mail: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share