You are on page 1of 82

Table of Contents

1 INTRODUCTION ...................................................................................................................................................7
1.1 Fundamentals ...............................................................................................................................................7
1.2 A Brief CAPWAP History ..............................................................................................................................7
1.3 Problem Statement .......................................................................................................................................8
1.4 Signal Matching – Step-by-Step ...................................................................................................................8
1.5 Resources ....................................................................................................................................................9
1.6 Limitations ....................................................................................................................................................9
1.7 Additional Considerations ........................................................................................................................... 10
1.8 The Question of Uniqueness and iCAP®.................................................................................................... 10
1.9 References ................................................................................................................................................. 11
2 THE CAPWAP PILE MODEL .............................................................................................................................. 12
2.1 Introduction ................................................................................................................................................. 12
2.2 Pile Segments ............................................................................................................................................ 13
2.3 Pile Properties ............................................................................................................................................ 13
2.4 Wave Propagation and Pile Variables ........................................................................................................ 14
2.5 Pile Slacks at Splices or Cracks ................................................................................................................. 16
2.5.1 The Standard Slack Model, Relative Displacement Limited ............................................................... 16
2.5.2 The Second Slack Model, Force Limited ........................................................................................... 16
2.5.3 Working with Slacks ........................................................................................................................... 17
2.6 Pile Damping .............................................................................................................................................. 19
2.7 Non-uniform Piles and Impedance Variations ............................................................................................. 19
2.8 Overall Wave Speed ................................................................................................................................... 20
2.9 Variable Time Increment............................................................................................................................. 20
3 THE CAPWAP SOIL MODEL .............................................................................................................................. 23
3.1 Basic Relationships .................................................................................................................................... 23
3.2 The Static Soil Resistance, Rsk ................................................................................................................... 24
3.2.1 The Unloading Limit, UN .................................................................................................................... 24
3.2.2 Loading and Unloading Quakes, QS, QT, CS, CT ............................................................................. 25
3.2.3 The Reloading Level, LS and LT ....................................................................................................... 26
3.2.4 The Toe Gap, TG ............................................................................................................................... 27
3.3 Dynamic Soil Resistance or Soil Damping.................................................................................................. 27
3.3.1 Smith Damping, SS, ST ..................................................................................................................... 28
3.3.2 Case Damping, JS, JT....................................................................................................................... 28
3.3.3 Damping Options, SO, OP ................................................................................................................. 28
3.3.4 Plugs (Acceleration Dependent Resistance), PS, PL ........................................................................ 29

CAPWAP Background Report Version 2014 1


3.3.5 Radiation Damping, SK, BT ............................................................................................................... 29
4 OTHER CAPWAP CALCULATIONS ................................................................................................................... 31
4.1 Match Quantity and Reflections at the Pile Top .......................................................................................... 31
4.2 Pile Toe Reflections .................................................................................................................................... 31
4.3 The Residual Stress Analysis (RSA) ....................................................................................................... 32
4.4 The Multiple Blow Analysis (MBA) ........................................................................................................... 33
4.5 Blow Count Calculation ............................................................................................................................ 34
4.6 Match Quality Assessment ....................................................................................................................... 35
4.7 Interpolation............................................................................................................................................... 36
4.8 PEBWAP .................................................................................................................................................... 36
4.9 Static Analysis ........................................................................................................................................... 38
4.9.1 Analysis ............................................................................................................................................. 38
4.9.2 Limitations .......................................................................................................................................... 38
4.9.3 Static Analysis Options ...................................................................................................................... 39
4.10 Data Adjustments ...................................................................................................................................... 41
4.11 Automatic Analysis Options ..................................................................................................................... 43
4.12 CAPWEAP ................................................................................................................................................. 44
5 CAPWAP UNKNOWNS AND SENSITIVITY ....................................................................................................... 46
5.1 CAPWAP Unknowns ................................................................................................................................. 46
5.2 Sensitivity Study - Example Description ................................................................................................. 46
5.3 Resistance Distribution Variation ............................................................................................................. 47
1.1.1.1 Figure 5.2.1 Final wave up match - US units............................................................................ 47
5.4 Shaft Damping Variation .......................................................................................................................... 48
5.5 Toe Damping Variation ............................................................................................................................. 49
5.6 Shaft Quake Variation .............................................................................................................................. 50
5.7 Toe Quake Variation ................................................................................................................................. 51
5.8 Unloading Limit Variation ......................................................................................................................... 51
5.9 Toe Gap Variation ..................................................................................................................................... 52
5.10 Unloading Shaft Quake Variation ............................................................................................................ 53
5.11 Unloading Toe Quake Variation ............................................................................................................... 53
5.12 Toe Plug Variation .................................................................................................................................... 54
5.13 Shaft Plug Variation .................................................................................................................................. 55
5.14 Radiation Damping Variation ................................................................................................................... 56
5.15 Toe Damping Option Variation ................................................................................................................. 58
5.16 Shaft Damping Option Variation .............................................................................................................. 59
5.17 Help ............................................................................................................................................................ 60
6 CAPWAP PROCEDURE ..................................................................................................................................... 61
6.1 Introduction ................................................................................................................................................. 61

CAPWAP Background Report Version 2014 2


6.2 Unknowns ................................................................................................................................................... 61
6.3 Analysis Steps ............................................................................................................................................ 61
6.4 Record Selection ........................................................................................................................................ 61
6.5 Data Adjustment ......................................................................................................................................... 63
6.6 Pile Model ................................................................................................................................................... 63
6.7 Signal Matching .......................................................................................................................................... 64
Example 1. Lack of Resistance .................................................................................................................... 64
6.8 Signal Matching Using Soil Model Extensions ............................................................................................ 68
6.8.1 Unloading Parameters CS, CT, UN ................................................................................................... 68
6.8.2 Reloading Levels, LS, LT .................................................................................................................. 68
6.8.3 Toe Gap, TG ...................................................................................................................................... 69
6.8.4 Damping Options OP, SO .................................................................................................................. 70
6.8.5 Radiation Damping, SK, BT ............................................................................................................... 70
6.8.6 Plug mass at toe, PL......................................................................................................................... 71
6.8.7 Extra Pile Toes................................................................................................................................... 71
6.9 Signal Matching with Non-Standard Pile Model Extensions ....................................................................... 72
6.9.1 Added Impedance and/or Shaft Soil Plug .......................................................................................... 72
6.9.2 Compression and Tension Slacks...................................................................................................... 72
6.9.3 Residual Stress Analysis, RSA .......................................................................................................... 74
6.9.4 Pile Damping, PI ................................................................................................................................ 74
6.9.5 Variable Time Increment and/or Wave Speed ................................................................................... 74
6.10 Additional Analysis Precautions .................................................................................................................. 76
7 CAPWAP LIMITATIONS AND APPLICATION RECOMMENDATIONS .............................................................. 77
7.1 CAPWAP (Dynamic Pile Testing) Limitations ............................................................................................. 77
7.1.1 Underprediction Due to Lack of Hammer Energy............................................................................... 77
7.1.1.1 Mitigation by the superposition of EOD and BOR method......................................................... 77
7.1.1.2 Underprediction mitigation by the superposition of BOR and EOR method .............................. 77
7.1.1.3 Precautions for the use of superposition ................................................................................... 78
7.1.2 Underprediction Due to Lack of Setup for EOD or Early BOR Tests .................................................. 78
7.1.3 Underprediction due to Plugging in Sands ......................................................................................... 78
7.1.4 Overprediction Due to High Energies ................................................................................................. 79
7.1.5 Overprediction Due to End Bearing Relaxation.................................................................................. 79
7.1.6 Overprediction in Highly Plastic Soils ................................................................................................. 79
7.1.7 Overprediction Due to Soil Column Adhesion in Cohesive Soils ....................................................... 79
7.1.8 Underprediction of Settlements .......................................................................................................... 79
7.1.9 Errors in Calculated Resistance Distribution and Uplift Capacity ....................................................... 80
7.1.10 The Uniqueness Question ................................................................................................................. 80
7.2 Comparing Apples with Oranges ................................................................................................................ 80

CAPWAP Background Report Version 2014 3


7.3 Additional Design Considerations ............................................................................................................... 80
7.4 Factors of Safety ........................................................................................................................................ 80
APPENDIX A: REFERENCES ..................................................................................................................................... 82

CAPWAP Background Report Version 2014 4


Table of Figures

Figure 2.1.1 The wave equation .................................................................................................................................. 12


Figure 2.2.1 Determination of pile model .................................................................................................................... 13
Figure 2.5.1 Two steel follower on concrete pile example (see Rausche, et al. 2006 DFI) ......................................... 18
Figure 2.5.2 Force-velocity record and CAPWAP impedance model for the situation depicted in Figure 2.5.1 .......... 18
Figure 2.5.3 Wave Up match for the situation depicted in Figures 2.5.1 and 2.5.2 ..................................................... 19
Figure 2.9.1 Example record of a slow wave return which may be caused by tension cracks .................................... 22
Figure 3.1.1 The CAPWAP soil resistance model is based on Smith with modifications ............................................ 23
Figure 3.2.1 Options for multiple toe resistance modeling (A) combining at the bottom and (B) individual toe
resistance effects at different pile segments ................................................................................................................ 24
Figure 3.2.2 The basic resistance force vs pile displacement model by Smith ........................................................... 26
Figure 3.2.3 The CAPWAP modified shaft resistance vs pile segment displacement model ...................................... 26
Figure 3.2.4 The CAPWAP modified toe resistance vs pile toe displacement model .................................................. 27
Figure 3.3.1 The complete CAPWAP soil model for the toe and for two shaft resistance segments near the pile toe.
..................................................................................................................................................................................... 30
Figure 4.6.1 Match quality time periods with a wave-up option ................................................................................... 35
Figure 4.8.1 Example of a PEBWAP result ................................................................................................................. 37
Figure 4.9.1 Extrapolation option (a) strain hardening, (b) strain softening ................................................................. 41
Figure 4.9.2 Calculated and smoothened elasto-plastic static load-displacement curve............................................. 41
Figure 4.10.1 Effect of A34 adjustment ....................................................................................................................... 42
Figure 4.10.2 Improper VT adjustment (top) and correct adjustment .......................................................................... 43
Figure 5.2.1 Final wave up match - US units .............................................................................................................. 47
Figure 5.3.1 80 kips (356 kN) taken from the toe and added to element 5 as friction ................................................. 48
Figure 5.3.2 16.9 kips (75.2 kN) taken from element 5 and added to the toe resistance ............................................ 48
Figure 5.4.1 Smith damping value increase ................................................................................................................ 49
Figure 5.4.2 Smith damping value decrease ............................................................................................................... 49
Figure 5.5.1 Smith toe damping value increase .......................................................................................................... 49
Figure 5.5.2 Smith toe damping value decrease ......................................................................................................... 50
Figure 5.6.1 Skin quake increase ................................................................................................................................ 50
Figure 5.6.2 Skin quake decrease............................................................................................................................... 50
Figure 5.7.1 Toe quake increase................................................................................................................................. 51
Figure 5.7.2 Toe quake decrease ............................................................................................................................... 51
Figure 5.8.1 Unloading limit decrease ......................................................................................................................... 52
Figure 5.8.2 Unloading limit increase .......................................................................................................................... 52
Figure 5.9.1 Small toe gap increase............................................................................................................................ 53
Figure 5.9.2 Large toe gap increase ........................................................................................................................... 53
Figure 5.10.1 CS decrease (in general CS should not be lower than 0.3) .................................................................. 53
Figure 5.11.1 CT decrease ......................................................................................................................................... 54
Figure 5.11.2 CT increase ........................................................................................................................................... 54
Figure 5.12.1 Plug weight increase ............................................................................................................................. 55
Figure 5.12.2 Plug weight decrease ............................................................................................................................ 55
Figure 5.13.1 PS = 0.2 kips (0.9kN) - segment 2 to toe .............................................................................................. 56
Figure 5.13.2 PS = 0.2 kips (0.9 kN) - lower half of pile .............................................................................................. 56
Figure 5.14.1 SK = 0.35 .............................................................................................................................................. 57
Figure 5.14.2 SK = 0.70 .............................................................................................................................................. 57
Figure 5.14.3 BT = 1.0 ................................................................................................................................................ 57
Figure 5.14.4 BT = 5.0 ................................................................................................................................................ 58
Figure 5.15.1 OP = 0 ................................................................................................................................................... 58
Figure 5.15.2 OP = 1 ................................................................................................................................................... 58
Figure 5.16.1 SO = 1 ................................................................................................................................................... 59
Figure 5.16.2 SO = 2 ................................................................................................................................................... 59
Figure 5.17.1 The HC Help Screen ............................................................................................................................. 60
Figure 5.17.2 The HR Help Screen ............................................................................................................................. 60
Figure 6.7.1 Force match showing lack of static resistance at top 3 segments ........................................................... 64
Figure 6.7.2 Excessive static resistance ..................................................................................................................... 65
Figure 6.7.3 Force match showing low resistance at segments 4 to 6 and high resistance below that ....................... 65
Figure 6.7.4 Force match showing good shaft resistance match, but low total static resistance ................................. 65
Figure 6.7.5 Force match showing excessive static resistance ................................................................................... 66
Figure 6.7.6 Force match showing evidence of insufficient damping .......................................................................... 66
Figure 6.7.7 Force match indicating need for non-proportional (added) damping, increased damping multiplier, added
impedance or plug on shaft .......................................................................................................................................... 67

CAPWAP Background Report Version 2014 5


Figure 6.7.8 Force match showing potential need for (unloading) quake decrease .................................................... 67
Figure 6.7.9 Force match showing the need for toe quake increase to delay toe resistance effect ............................ 67
Figure 6.8.1 Force match with low computed force due to either high UN value or low CS or CT .............................. 68
Figure 6.8.2 Reloading along slope given by loading quake (low reloading level) ...................................................... 69
Figure 6.8.3 Reloading with high reloading level ......................................................................................................... 69
Figure 6.8.4 Example of a load-set curve showing soil resistance decrease with downward pile movement ............. 70
Figure 6.8.5 Example of an upward traveling wave which displays typical unloading characteristics ......................... 71
Figure 6.8.6 Example of a match where a plug mass at the toe, PL, could improve the match .................................. 71
Figure 6.9.1 Record example of a pile with tension cracks ......................................................................................... 73
Figure 6.9.2 Record example of a potential compression slack plus evidence of a tension slack at a splice location 74
Figure 6.9.3 Pile with strong impedance reduction...................................................................................................... 75
Figure 6.9.4 Pile with lesser impedance reduction ...................................................................................................... 75
Figure 7.1.1 Synthetic example of superposition of several restrike results ................................................................ 78

CAPWAP Background Report Version 2014 6


1 INTRODUCTION

1.1 Fundamentals
Measurements taken with the Pile Driving Analyzer® on pile, usually near the pile top, during an impact can
and should be evaluated by the Case Method. This method allows for a quick and simple calculation of soil
resistance, integrity and stress values in closed form which means that limiting assumptions such as pile
uniformity have to be made.

Dynamic pile analysis by the wave equation, using GRLWEAP, allows for the calculation of stresses and pile
motions of sometimes very complex situations. This numerical analysis requires that hammer, pile and soil are
modeled with a number of segments which are represented by their masses, stiffness values and damping
factors.

CAPWAP® combines the measurements of the Case Method with a wave equation type numerical analysis.
Since the pile top motion and forces have been measured it is then not necessary to model hammer and
driving system. Using the measurements as the input function, the soil model can be adjusted so that the
computed response matches the measured response, and thus also the process is referred to as “signal
matching”. Such a numerical analysis approach of the measurements yields a complete set of results for a
complex pile and soil systems. The limiting assumptions in this case are the simplificatio ns of the basic soil
model. Thus, CAPWAP overcomes the limitations of the Case Method (and so does its younger and not quite
as capable cousin, iCAP®, as described below) and in addition to providing a complete set of pile variables
(forces, velocities, displacements, stresses, accelerations) also results in the most likely soil resistance
distribution and its dynamic (damping) components.

Considering that Case Method and GRLWEAP are foundations for CAPWAP, the reader of this Background
Report would do well with studying the applicable manuals of those two Methods as it would greatly help with
the understanding of the CAPWAP models and input parameters.

1.2 A Brief CAPWAP History


A number of developments leading to today’s software are noteworthy. The firs t working signal matching
program was developed in the late 1960s at Case Institute of Technology under a grant by the Ohio
Department of Transportation. This program, developed on a main frame computer, used a Smith -type lumped
mass pile model. Documented by Rausche, 1970 and by Rausche et al., 1972, it ran automatically and
provided reasonable results for the relatively short piles for which force and velocity data was available.

Further developments were necessitated as computers became either more ine xpensive or more powerful or
both and as piles tested became more varied (concrete, timber, composite), larger and longer. First the basic
Smith soil model had to be expanded, most notably with independent unloading parameters and different
damping approaches. Then the limited capabilities of early minicomputers and microcomputers required that
the program could also be operated in an interactive manner. Next, for long piles it was beneficial to move
from the lumped mass model to a pile model based on the Method of Characteristics (Rausche, 1983). This
program was referred to as CAPWAPC, however, in 1993 when additional developments were presented to
the PDA Users, the program and its successors were then again referred to as CAPWAP even though they
continued to use the characteristics approach.

With the arrival of ever more powerful microcomputers, it became feasible to reinstate the automatic features
of the earlier versions and add to those. A program that provided both interactive and a powerful automati c
feature was developed in the latter half of the 1980s. Finally, extensive changes were necessitated by the
arrival of the Windows Operating System. Taking advantage of the new graphics and expanded memory
capabilities, CAPWAP grew further and became more and more user friendly with more automatic features and
expanded analysis capabilities (Radiation Damping, Residual Stress Analysis, Blow Count Matching, Plug
Models etc.) while the basic analysis model remained unchanged (Rausche, et al., 2000).

Since 2011 Pile Dynamics has responded to the general request for a reliable signal matching program which
is fast enough to be performed by the Pile Driving Analyzer in real time, i.e., during monitoring. This method,
called iCAP, while not as powerful and accurate as CAPWAP is expected to gradually replace the Case
Method as the preferred method for calculating soil resistance during pile driving monitoring (Likins et al.,
2012). Further discussion on iCAP may be found in Section 1.8.

CAPWAP Background Report Version 2014 7


1.3 Problem Statement
In the dynamic analysis of piles (including drilled shafts or bored piles) under an impact or transient dynamic
load there are four sets of unknowns: internal pile forces, pile motions, soil resistance forces and soil motions.
From measurements both force and velocity, usually near the pile head, are accurately known (if not, then a
sophisticated analysis would not help), however, the static and dynamic soil resistance forces plus all pile
forces and motions along the pile length and soil motions are unknown. Unfortunately, in general, direct
calculation of the complete soil model from measured pile top force and velocity is not practical. (An exception
is a pile with only toe bearing and it can be analyzed in the PEBWAP routine of CAPWAP). However, we can
analyze a fairly complex system with either the top force or the top velocity or their average (which is the force
in the downward wave) prescribed at the model top and subject to soil resistance forces based on an assumed
soil model. The other, unused measured quantity, velocity, force or wave up, is then plotted and compared
against the equivalent, computed variable. Differences between the measured and c omputed curves lead the
experienced engineer to reflect on the differences between the actual soil behavior and the assumed set of
soil parameters. Modification of these assumed soil parameters should lead to a better match in a subsequent
iteration. CAPWAP (CAse Pile Wave Analysis Program) was written to facilitate this procedure which is also
referred to as a "System Identification", “Signal Matching” or “Reverse Analysis”.

Soil reaction forces are passive, i.e., they change only when the associated pile segments change
displacement due to the traveling stress wave (pile segment has a velocity). Similar to GRLWEAP, CAPWAP
assumes that the soil reaction consists of elasto-plastic and linear viscous components. The soil’s static
reaction relates to the pile segment displacement and the soil damping relates to the pile segment velocity. In
this way, the soil model has at each point three unknowns: the ultimate static resistance, the quake or elastic
soil deformation, and a damping constant. The CAPWAP analysis is completed when a best possible match
between measured and computed variables has been obtained. Then these three unknowns have been
determined for a number of discrete points along the pile and at the pile toe.

When the match is unsatisfactory, the process of iteratively changing the soil resistance parameters and
computing the pile top variable is repeated. To compute the pile top variable for comparison with the measured
one, algorithms are followed that allow for a step-by-step computation of all pile variables along pile length and
over time. Many of the important equations that are used in this process will be listed in the following chapters.
Also the parameters of the soil model will be described both mathematically and physically and thei r effect on
the “Match” will be discussed. This should help the user with making correct input choices for an expedited
matching process.

1.4 Signal Matching – Step-by-Step


The basic signal matching procedure can be summarized as follows:

i. Select a force/velocity record of a significant hammer blow for analysis (e.g., last blow of installation or
early blow of a restrike or dynamic load test)
ii. Set up a mechanical model of the pile by dividing the pile into a series of short “pile segments”, with the
top of the model being the point where the force and velocity were measured.
iii. Set up a mechanical model of the soil by representing distributed soil resistance into concentrated soil
resistance forces at selected pile segments. In CAPWAP the soil model is based on an expanded
Smith model.
iv. Make a first guess at a set of soil resistance parameters (e.g., ultimate capacities, quakes and damping
factors for all soil segments. Actually the CAPWAP program automatically makes this first guess.)
v. Analyze the pile and soil system by prescribing at the pile top one of three possible, measured
quantities: force, velocity or force in the wave-down. Calculate the complementary quantity: velocity,
force or force in wave-up.
vi. Compare the calculated, complementary top quantity with the corresponding measured quantity and
calculate a “match quality” (MQ) by summing up differences between the two quantities over time.
vii. Evaluate the differences between computed and measured curves and recommend changes of soil
resistance parameters to improve the match. Modify the parameters and return to v.
viii. After calculating in a first cycle the static resistance parameters under the assumptions of certain
damping and quake factors, vary these additional dynamic soil resistance parameter s systematically,
always returning to v. for reanalysis.
ix. When it is not possible to further improve the match quality, the static soil resistance has been
determined and output can be made which will, in addition to summarizing the soil resistance
parameters and showing the pile model, also involve performing a t-z/q-z analysis to produce a
simulated static load test.

CAPWAP Background Report Version 2014 8


1.5 Resources
This document is only one part of the CAPWAP documentation. It describes the mathematical models and soil
resistance parameters that form the basis on which CAPWAP makes its calculations and is therefore called a
“Background Report”. This report will not provide the reader with “How To” information. Help and Examples are
instead provided on line and distributed along with the software package.

Again, it should be emphasized that the information contained in the CAPWAP Background Report can only be
understood if the reader has a solid knowledge of:

• Measurement and Data Processing by the Pile Driving Analyzer


• One-dimensional wave mechanics as applicable to pile analysis
• Case Method analysis of dynamic pile testing records
• Ideally also the GRLWEAP wave equation analysis

In fact, one part of the CAPWAP analysis procedure is the execution of the PDA software program during data
selection and data quality assessment and, if necessary, data adjustment. It is therefore essential that the user
first is familiar with the PDA Manual before attempting CAPWAP. Much of CAPWAP, in particular its soil
model, is also very similar to the GRLWEAP wave equation approach and an understanding of that program
would be very helpful.

The user should undergo formal training and is encouraged to contact Pile Dynamics (PDI) if questions during
the program execution arise or if the program does not appear to perform satisfactorily. Furthermore, the
CAPWAP user should attend the PDA/CAPWAP Workshops, where proven, new and improved approaches for
record interpretation are discussed. Only such preparation, training and continued education can advan ce the
user’s skill in the art of "CAPWAPing", improve the accuracy of predictions and, inadvertently, make avoidable
mistakes.

1.6 Limitations
The success of the CAPWAP analysis and its ability to accurately predict the static behavior of a pile therefore
depends on several assumptions, including but not limited to:

i. good quality data was measured in the field;


ii. the soil resistance model is realistic and able to separate the static from the dynamic resistance
components;
iii. the pile has been properly represented;
iv. the hammer energy was sufficient to result in a sufficient permanent displacement of the pile;
v. the signal match is satisfactory;
vi. if the long term static behavior is to be assessed, the dynamic pile test was performed during a
restrike with sufficient waiting time after pile installation (CAPWAP calculates the pile capacity at the
time of testing).

The CAPWAP soil model is in many parts identical to that developed by Smith (1950, 1960) and this model
has been extensively tested on many projects. However, this Smith model is relatively crude considering the
complexity and variability of soil properties not only around the world but even at the same site. The
shortcomings of this model (which make it attractive because it is a manageable o ne) are, for example,

i. the rate sensitivity (generally called the damping behavior) is linearly related to the pile velocity so that
the static resistance to driving (SRD) is the same under very slow or very rapid loadings,
ii. that the soil surrounding the pile is considered at rest, and
iii. the pile and soil stresses are zero immediately prior to impact.

Extensions to the basic Smith model (see also Rausche et al., 2010) had to be provided in CAPWAP (a) to
obtain reasonably good matches of the measured and computed pile top variables and (b) to overcome the
limitations ii and iii above. Yet, limitation i, is the most fundamental assumption of the CAPWAP approach. It
may be violated in a very few instances of very plastic cohesive soils.

Admittedly, these basic assumptions and conditions make for a difficult analysis environment (see also, for
example, Likins et al., 2008, Rausche et al., 2008) and that is why PDI strongly recommends that the user
attempts to fully understand the physical limitations of all soil model details before accepting the final
CAPWAP results. Fortunately, comparisons between static load test capacities (see for example, Likins et al.

CAPWAP Background Report Version 2014 9


1996 and Likins et al., 2004) and those calculated by CAPWAP are very good even considering the rat her
varied conditions under which the results were obtained.

1.7 Additional Considerations


As mentioned earlier, the CAPWAP results are based on the "best possible match" between a computed pile
top variable, say the pile top force, and its measured equivalent. Once the match has been optimized, the
analysis is finished. It could be argued that the results of the analysis should also match what has been
determined by soil borings or in-situ tests. For example, knowing a Cone Penetrometer profile might indicate a
different resistance distribution or would suggest a different damping behavior. However, imposing certain soil
resistance parameters, such as the static resistance along the length of the pile would not lead to a best
possible match. This may be confusing, but is possibly the result of (among other reasons):

 the dynamic action of the pile driving process on the soil (e.g., remolding, densification or loss of
density, pore water pressure effects);
 the variability of the site;
 uncertainties in the effect of the soil investigation process (including sampling methods);
 uncertainties in the assessment of soil resistance parameters based on static formulas and soil profile
information. For example, scaling effects (from a small penetrometer to a large pile) a re known to put
into question the static analysis predictions (that’s why we are testing);
 variations in overburden (excavation or surcharge) and water table between the time of soil exploration
and pile driving;
 the time effects (setup or relaxation) of the soil taking place after installation;
 differences in the soil resistance behavior during static and dynamic load applications; plugging, for
example, may or may not occur during the dynamic test but may happen during the static test. The
reverse situation, a plug moving with the pile during driving and slipping under static loads, is also
possible.

Thus, because of these differences between static capacity predictions and actual performance, imposing on
the CAPWAP solution what is expected from a soil profile and associated static geotechnical analysis would in
effect make the test useless, and usually not result in a best match. However, after CAPWAP has provided a
set of soil resistance parameters, it is the engineer’s duty to check the significance o f all information collected.
The CAPWAP result is just one of the important results that make up the basis of a geotechnical foundation
design and construction control process. (Similarly, even a static load test should be evaluated in light of
geotechnical conditions or changes – e.g. downdrag conditions or compressible soil layers).

Even though it is evident that the simplified Case Method has more severe limitations than CAPWAP and
needs as an input an assumed Case damping factor, it is sometimes attem pted to “proof” that a certain Case
Method capacity value (or Case damping factor, JC) is the correct one. In general, the best match cannot be
achieved by CAPWAP in this way and trying to force a match with a certain Case Method as the CAPWAP
calculated capacity would make the CAPWAP analysis superfluous.

1.8 The Question of Uniqueness and iCAP®


Purists would like to get one result and only one result. More pragmatic engineers who realize the complexity
of the deep foundation test would like to know upper and lower bounds of a solution. All inverse analysis
procedures, however, have a somewhat tainted reputation in that they may give a number of solutions which
match the underlying set of measurements.

As for CAPWAP, theoretically, each point in time of the two measured pile top variables force and velocity
provides information which supports the determination of one soil resistance parameter. So for a 1000 point
record length there are 2000 pieces of information while the number of unknown soil model paramete rs will be
less than 60 for a pile of about 40 m embedment. So that is one argument that the CAPWAP solution is
unique.

Another argument for uniqueness is that for one set of measurements, and one pile model, and one defined
set of soil model parameters, and one objective function (e.g. MQ), there is only one set of soil model
parameters that will yield the absolutely best (lowest) MQ value. From a practical view, it is not realistic to
search a nearly infinite array of possible soil model parameter combinations to locate this absolutely best and
hence unique solution. But neither is such an exhaustive search really needed in practice.

CAPWAP generally yields a unique capacity value, i.e., the best match and matches that are giving nearly
identical match quality values, generally predict the same capacity value. However there is no denying that, for

CAPWAP Background Report Version 2014 10


example, end bearing and shaft resistance near the toe are not necessarily uniquely determined (particularly
when shaft and toe quakes are similar) and the less the pile toe moves (insufficient energy) the less clearly
can the differentiation be made. Clearly, the lower the number of parameters that a model has, the more
uniquely defined is its result, however, wrong that result may be.

PDI has responded to the requirement of obtaining only one result and to do that with a signal matching
procedure which has become a requirement by various building codes or specifications. This is done with the
iCAP Method (Likins et al, 2012) which is a completely automatic si gnal matching procedure (no user
interaction) which is fast enough that it can be done in real time, often for all hammer blows recorded. The
iCAP results are not considered as accurate as those determined by a thorough CAPWAP analysis, but they
compare very well with CAPWAP for normal pile and soil conditions. Best of all, they are unique and available
at the end of the test and indicate both shaft resistance and end bearing components.

1.9 References
PDI’s web site offers a wide variety of publications which cover many different aspects of dynamic pile testing
applications and CAPWAP analyses. The appendix shows references that are directly referred to in this
document.

CAPWAP Background Report Version 2014 11


2 THE CAPWAP PILE MODEL

2.1 Introduction
CAPWAP’s numerical analysis is based on the so-called method of characteristics which solves the famous,
linear one-dimensional wave equation which is shown along with a very general solution in the following slide
(Figure 2.1.1).

Figure 2.1.1 The wave equation

For realism and generality, CAPWAP applies the solution which involves a downward traveling wave, f, and an
upward traveling wave, g, to a series of short pile segments (each segment is uniform, but segments may vary
along the length of the pile if the pile is non-uniform) representing the whole pile. The basic principle is shown
in the next slide. The center segment, numbered i, has a downward wave, F dti starting to travel down the
segment and arrives at the bottom of that segment a time increment Δt later as F dbi . These two downward
waves are of the same magnitude. For the upward wave however, since a total resistance force acts against
the bottom of segment i, the arriving upward wave at the top, F uti equals F ubi + ½Rti . The other half of that
resistance force modifies the downward wave at the next lower segment. More detailed equations will be
shown below. Note that at segment interfaces where either the cross sectional area A i or the the pile material
properties E i or ρi change, reflections will occur which affect the force magnitude in the downward wave at the
top of the next lower segment or the force in the upward wave at the bottom of the segment above. All
segments have to have the same time wave travel time so that at the same time the individual waves arrive at
their next boundary at the same time. Thus, the computation time increment, Δt, must be the same for all
segments and if the wave speed c i is different in some of the segments then their length increments also has
to be different, i.e., ΔL i = ci Δt. Obviously the impedance Z i is not only depending on the cross sectional area
Ai but also on the material properties, i.e., Z i = Ei Ai / ci = ρ i c i Ai = Ai (Ei / ρi )0.5 .

The program calculates the wave down and wave up values in each segment at times when downward
(upward) waves arrive at the bottom (top) of the segments. At that point reflections are calculated from the
relative properties of neighboring segments and the resistance forces. Since superposition of waves is
possible according to the underlying one-dimensional differential equation (the “wave equation”), total forces
and velocities for each pile segment can then be directly calculated from the sum (difference) of the downward
and upward wave values. The displacement of each segment is calculated by Euler integration from the
velocity.

CAPWAP Background Report Version 2014 12


2.2 Pile Segments
The Np pile segments (Figure 2.2.1) should be chosen not much longer than ΔL = 1 m (3.3 ft). For steel with a
wave speed of about 5100 m/s the time increment is then 1.0/5100 or somewhat less than 0.2 ms. For
cushioned hammer blows this is a time increment that is consistent with the higher frequency components of
the pile top measurements. However, for uncushioned impacts, the frequencies may be higher and it is then
advisable to use smaller segments by, for example, doubling the N p value. Another case where it is advisable
to use shorter segments is for piles where nonuniformities (e.g. connectors) happen a certain regular
distances and making the segment lengths such that they occur where the cross sectional changes happen,
would make for a more accurate model otherwise the change of cross section would cause average properties
in neighboring segments.

Figure 2.2.1 Determination of pile model

2.3 Pile Properties


Cross sectional area, elastic modulus, specific mass (usually expressed as a specific weight times the
gravitational acceleration, g = 9.81 m/s 2), and pile perimeter are the four quantities that make up the “Pile
Profile”. The pile perimeter is only needed for unit resistance calculations in the final output. The other three
quantities define the standard dynamic pile model. For uniform piles, the pile top properties are sufficient to
define a pile profile. For non-uniform piles, the pile profile has to specify, as a function of depth, where any
one of the four basic quantities changes. Two additional pile properties are the pile toe area (for the
calculation of unit end bearing) and the overall wave speed, if it differs from the wave speed calculated from
elastic modulus and specific weight. In the PDA software, the former corresponds to WC (calculated wave
speed) and the latter to WS (elastic wave speed).

For given pile properties of pile segment i, E i , ρi and the wave speed, c i , may be averaged over the segment
length ΔLi if the properties change within the segment. The wave speed is calculated from the two fundamental
material properties as

ci = (Ei / ρi)½ (2.1)


As we have seen, each segment, i, has a length ΔL i , such that its wave travel time, Δt i , equals the analysis
time increment, Δt.

ΔLi = Δt ci (2.2)

The sum of all segment lengths ΔL i (with i equal to 1, 2, …, N p) equals the pile length below sensors, L S, and
Np(Δt) is equal to the total wave travel time L S/c O where c O is the overall wave speed. Since each pile
segment is uniform and linearly elastic, the magnitude of the force in a downward traveling wave, F ij, at time j
at the top of a segment i, is the same as the wave arriving at the bottom of the same segment at time j + 1
(i.e., one time increment Δt later).

CAPWAP Background Report Version 2014 13


CAPWAP calculates the pile velocities and displacements of the individual model segments according to the
method of characteristics (Rausche, 1983). Each segment has an impedance

Zi = (Ei Ai )/ci (2.3a)

Considering Eq. 2.1 the impedance can also be calculated from

Zi = Ai (Ei ρi )0.5 (2.3b)

Which more clearly expresses the fact that the impedance is a function of both pile cross section (size) and
material properties (quality).

For calculation of the load - set curve of a pile under a static load, the pile stiffness, k i is also needed and can
be calculated using the following equation

k i = Ei Ai /ΔLi (2.4a)

which after multiplying numerator and denominator by c i yields

k i = Ei Ai /ci (c i /ΔL i ) = Z i /Δt (2.4b)

For completeness sake it may be mentioned that the mass of a segment is

m i = ρi Ai ΔLi = (E/ci 2) Ai ΔLi = Z i Δt (2.5)

and multiplying Eq. 2.4b with Eq. 2.5, the pile impedance can also be written as

Zi = (k i mi )½ (2.6)

Finally, dividing Eq. 2.5 by Eq. 2.4b one obtains

Δt = (m i / ki )½ (2.7)

The above formulas can be very helpful. For example, if for one reason or another it is necessary to modify the
impedance of a segment, i.e. make it different from the original pile profile, then assuming that the pile
properties E i and ρi did not change, Eq. 2.3 allows for calculating the new area A i and therefore also the new
volume A i ΔLi of the modified segment. Stiffness and mass of the modified segment are then obtained from Eq.
2.4b and 2.5.

2.4 Wave Propagation and Pile Variables


According to basic wave mechanics, the forces and velocities in the upward and downward traveling waves
are proportional. Thus, with F d (Fu) designating the force in a downward (upward) wave and v d (vu) standing for
the velocity in a downward (upward) wave, the force in the downward traveling wave is,

Fd = Zvd (2.8a)

Similarly, the force in the upward traveling wave is,

Fu = - Zvu (2.8b)

The second important set of equations states that, for linearly elastic rods, forces ( and velocities) at a point
can be calculated by superposition of the upward and downward waves. Thus,

F = Fd + Fu (2.9a)

and

v = vd + vu (2.9b)

Equations 2.8 and 2.9 lead directly to the forces and velocities in the upward and downward waves. Let us
label the pile segments with 1, 2, …, i, … N p beginning at the segment just underneath the point where the
measurements were taken. Thus, the force in the downward traveling wave F di (t) at segment i and time t can

CAPWAP Background Report Version 2014 14


be calculated from the total force, F i (t) and total velocity, v i (t), in the segment using:

Fdi (t) = [Fi (t) + Z i vi (t)]/2 (2.10a)

Similarly, the force at segment i in the upward traveling wave is,

Fui (t) = [Fi (t) - Zi vi (t)]/2 (2.10b)

With time being discretized, at any time index j (j is 0 when the analysis begins and then increases by 1 with
each time increment so that the time is j Δt at time increment j), both upwards and downwards traveling waves
are present in segment i. If neighboring segments have equal properties and if there is no resistance force
acting along these segments, then one can calculate the downward (upward) wave force at time j+1 in
segment i from the downward wave at time j in segment i-1 (i+1). Thus, for two neighboring segments of equal
impedance, the forces in these waves are

Fui,j+1 = Fui+1,j (2.11)

and

Fdi,j+1 = Fdi -1,j (2.12)

If the cross sectional impedance properties for segments i and i+1 are different, then the pile impedance, Z i
(see Eq. 2.10), has to be considered for reflections occurring at the interface between the segments of
different impedance. Let us define impedance ratios between segments i-1 and i and i+1 and i as follows.

Za,i = Zi /(Zi + Z i+1 ) (2.13)

Zb,i-1 = Zi /(Zi + Zi-1) (2.14)

Not only impedance changes but also the total soil resistance R k (static plus dynamic) at pile element i affects
the waves in the segments neighboring segment i. Considering these effects, the downward and upward wave
forces may be calculated from

Fui,j+1 = Za,i [2F ui+1,j - Fdi,j + Rk] + Zb,i+1 Fdi,j (2.15)

Fdi,j+1 = Zb,i [2F di-1,j – Fui,j - Rk] + Z a,i-1Fui,j (2.16)

In these equations, R k is the k th soil resistance acting at the bottom of pile segment i. Once the upward and
downward traveling wave forces are known, the associated velocities are simply

vui,j = -Fui,j / Zi (2.17)

vdi,j = Fdi,j / Zi (2.18)

The segment force and velocity can now be calculated from superposition as

Fi,j = Fui,j + Fdi,j (2.19)

vi,j = vui,j + vdi,j (2.20)

The pile segment velocity, v ij, is integrated to displacement, u ij, of segment i and at time step j from:

u i,j = ui,j-1 + ½ (vi,j-1 +vi,j) Δt (2.21)

Of course, the acceleration could be calculated by differentiation. However, the acceleration is not generally
useful in CAPWAP (in contrast to the GRLWEAP program which calcu lates acceleration which are then
integrated to velocity and displacements. The stresses in the pile are calculated by CAPWAP from segment
forces as averages over the segment cross sectional area.

σi,j = Fi,j / Ai (2.22)

CAPWAP Background Report Version 2014 15


2.5 Pile Slacks at Splices or Cracks
In GRLWEAP, the modeling of non-linearities such as occur at interfaces of splices or cracks or other
interfaces with slacks (a slack is defined as a distance over which neighboring segments can separate without
transmitting either a full tension or compression force). In CAPWAP’s characteristic model this is somewhat
more complicated and instead of defining the slack with a round-out distance and a slack distance (as in
GRLWEAP), the standard model works with a slack distance and a so-called slack efficiency (which defines
whether or not full or partial reflections occur). This model does not always provide the desired effect. For
example, a tension crack may open but the reinforcement still transmits force although with a low stiffness so
that the transmitted force is much smaller than would be for the uncracked section for the same segment
extension. The second force slack option only works with a limiting force value and does not consider an
efficiency. Standard slack models and force limiting slack models cannot be mixed in one analysis. In other
words if at one pile segment a force slack has been defined then all the other segments also are defined with
force slacks. Another restriction is that the CAPWAP pile model allows the user to include for a particular
segment to model only either a tension force or a compression force slack (deformation slacks of both types
can occur on the same segment).

As mentioned a crack or splice may cause complete reflection of a compressive wave if it is open or may
transmit compressive waves if it is closed. After the slack distance has been overcome, full wave transmission
resumes. Thus, a tension crack would transmit tension waves after it is fully extended.

2.5.1 The Standard Slack Model, Relative Displacement Limited


Putting this in a mathematical form, a slack may transmit either tension or compression waves after a tensile,
St, or compressive, S c, slack distance has been reached or exceeded by the relative displacement u r = ui+1 - u i
between the two segments i and i+1, neighboring the splice i. Thus, waves are transmitted only if; u r > Sc and
u r < St. The slack values S t and S c are called slack distances.

Unfortunately, this simple slack model does not represent reality sufficiently well to match actual crack or
splice behavior. The model was therefore modified to always allow some force to be transmitted across a
crack or splice, e.g., due to reinforcing, bending cracks extending over only part of the cross sect ion, or other
non-perfect separations. The partial wave transmission is expressed through a so -called slack efficiency, r.
The user specified slack efficiency, r, can take on values between 0 and 1. Depending on the magnitude of r,
the downward wave is at least partially reflected upwards and the upward wave is at least partially reflected
downwards, as long as the slack is open. With r = 0 the slack is permanently closed like for no splice at all. A
slack efficiency of 1 expresses that the slack is fully open whenever the separation between the sections is
less than the splice distance. This makes for a very harsh transition between full wave transmission and no
transmission. In the case of r between 0 and 1 and the slack being open the wave transmission fo r the upward
traveling wave is calculated as follows.

Fui,j = (-Fdi,j-1 + Rk/2) (r) + (F* ui,j)(1-r) (2.23)

The F*u i,j term is the upward calculated wave for an unspliced section (full wave transmission). The
corresponding downward wave is

Fdi+1,j = (-Fui+1,j-1 - Rk/2) (r) + (F* d+1,j)(1-r) (2.24)

Typical efficiency is relatively low (e.g. 0.1) but has to be determined by trial and error.

2.5.2 The Second Slack Model, Force Limited


In the second model, the compressive or tensile slack inputs, S c or St are interpreted as a maximum slack
force F s (either tensile or compressive depending on the specified slack type). The following equations allow
for open slacks with unlimited slack distance, but with a constant limiting force between the two neighboring
sections.

Fui,j = -Fdi,j-1 + Rk/2 + F s (2.25a)

and

Fdi,i+1,j = -Fui+1,j-1 - Rk/2 + F s (2.25b)

Note (a) that there is no efficiency used in this model and (b) that the reflection would be the same as in the

CAPWAP Background Report Version 2014 16


standard model with an efficiency equal to 1 and an unlimited slack distance, if F s were 0.

2.5.3 Working with Slacks


Modeling cracks or splices with slacks is not a simple task because (a) two models exist, (b) we don’t know
exactly where to put the slacks (compressive slacks are easier to locate than t ensile slacks because we can
see their reflection effect during the first 2L/c), (c) the standard tension slack model involves two unknowns
(slack distance and slack efficiency) and (d) even if we have found a decent match with one slack location and
set of parameters, it may be possible to find even better agreement with maybe more than one slack at
different locations with different slack distances, efficiencies or force values (particularly true for tensile
slacks).

Compression slacks are not very common; they have to represent a splice or crack which is open at the
beginning of a hammer blow. Finding the location at which segment they should be placed should, therefore,
be relatively straight forward. In general, compressive slacks of the standard model have to use very small
efficiencies and distances or the reflection would be very large. The second limiting force model is rarely used
for compressive slack modeling. It may be appropriate for a friction type splice (neighboring pile sections are
held together by a friction sleeve). They also would possibly be helpful to model a damage (reduced
impedance) in addition to a crack. In such a case, instead of using compressive slacks it may be simpler to try
a reduced impedance.

Tensile slacks are more common. They may be needed to represent, for example, a mechanical splice which
has some slack or the interface between the bottom of a follower (chaser) and the top of a pile. In those cases
it is easy to locate the tension slack and also its distance is relatively well defined (may be one mm in the
splice case and rather large or unlimited in the follower case). Then only the efficiency needs a trial and error
adjustment. It is more troublesome to locate and model a tension slack that should represent a cra ck in the
pile. A tension crack reveals itself only when a tension wave opens it up. This can happen for low resistance
situations during the time between L/c and 2L/c. For high resistance pile driving it may also happen during the
downward wave travel after 2L/c. Large quake, high end bearing situations may have crack openings occur
during both upward and upward wave travels and will only reveal themselves by either a cut -off tension
reflection and/or a much wider tension “valley” at 2L/c compared to the appearance of the inverted impact
wave width for the uncracked pile. To find a reasonable location in the easy driving situation, it may be
instructive to determine where maximum tension stresses occur (see the PDA Manual section on tension
stresses) for a clue as to where the most likely location of a tension crack would be (typically in the upper 1/3
of the pile length). It may be close to the top where the downward wave frequently has a reduced effect.

Occasionally it may be necessary to utilize several tension slacks to achieve an apparent slowing of the wave
propagation caused by the opening of several small tension cracks. That can be done with very small slack
distances and efficiencies at every second or third segment. It may be easier, however, to check whether a
reduced overall wave speed (see Section 2.7) helps; unfortunately, that would have an effect already at a time
when there is no tension present in the pile. In simple cases, a slightly increased pile damping (Section 2.5)
may do the trick. If these means are difficult to use or ineffective, then the variable time increment option of
the 2014 version can be tried (Section 2.8).

An example where both compressive and tensile slacks have to be utilized for a successful CAPWAP analysis
is the interface between a follower on the top of a concrete pile protected by a relatively soft wood cu shion. As
mentioned, often times it is easier and also acceptable to model such an interface with a reduced impedance
rather than with a compressive slack (see also Section 2.7. However, a tension slack will always be needed in
such a case. Note that if force type tension and compression slacks are needed then they have to be input for
neighboring segments, because they cannot occur on the same segment.

As an example, consider the situation of two followers driving a concrete pile ( Figure 2.5.1 and Rausche et al.,
2006). The CAPWAP result page pertaining to that case is shown in Figure 2.5.2. The two steel followers had
an identical cross section (but with top and bottom plates) driving a concr ete pile protected by a wood cushion
at its top. The force-velocity record (upper right hand corner) shows very clearly a compressive reflection
(temporarily force increase relative to the velocity) followed by a tensile reflection (temporarily velocity
increase relative to the force) from the point where the upper steel follower sits on the lower steel follower
connection. There follows a much larger tensile reflection from the wood cushion on the concrete pile. The
concrete pile is a hollow straight section on top of a more or less solid tapered section.

CAPWAP Background Report Version 2014 17


Figure 2.5.1 Two steel follower on concrete pile example
(see Rausche, et al. 2006 DFI)

Figure 2.5.2 Force-velocity record and CAPWAP impedance


model for the situation depicted in Figure 2.5.1

The model employed on average 0.5 m long segments for increased resolution of the complex geometry.
Actually, the steel segments and concrete segments were roughly 0.6 and 0.4 m long, because of the
differences in their wave speeds. For the steel follower connection a fairly good match was achieved with first
an increase of pile impedance and then a decrease back to nominal pile impedance. The impedance vs. pile
length variation is shown directly beneath the force and velocity record. The steel -wood-concrete connection
was represented by an unlimited tensile slack (99 mm) with 0.65 efficiency and a force limited compression
slack of 6.5 kN. For a best match, the location of the impedance changes had to be moved a slight distance
away from their actual position. Obviously this is a rather complex situation and it is not surprising that the
match was not exact (Figure 2.5.3).

CAPWAP Background Report Version 2014 18


Venice; Pile: Ven-PC14b-2SF; CONCRETE-DOUBLE STEEL FOLLOWER; Blow: 155 (Test: 19-Apr-2005 11:28:)
Beta Version - Pile Dynamics

800 2000
W up Msd kN
m/s
W up Cpt

1250

15 75 ms
0 500
4 L/c 15

-250

-1000
D is p la c e m e n t ( m m )
-800
Pile Imp

Figure 2.5.3 Wave Up match for the situation depicted in Figures 2.5.1 30.00
Load (kN) Shaft Re
and 2.5.2 Pile Top 25.00 Distribut
0 70 140 210 280 350
0.0 Bottom
20.00

k P a
15.00

2.6 Pile Damping 13.0 RU =


SF =
306 kN
184 kN
10.00
EB = 123 kN 5.00
No traveling wave propagates
26.0 forever. Internal pile material damping may be modeled by 0
0.00
computing the Dy = 39.0 mm
Dx = 63.9 mm
change of a wave and reducing the new wave by a specified fraction, j . Thus, p=
SET/Bl 26.0 mm
EB
80
39.0
160
F*ui,j = Fui,j - j p (Fui,j - Fui,j-1 ) (2.26a)

k N
SF
52.0 240

F*di,j = Fd,j - j p (Fdi,j - Fdi,j-1 ) 320 (2.26b) P


a
65.0 400
The * indicates the dampened wave value. The pile damping value, j p (PI is the corresponding input quantity)
is accessible both in the main CAPWAP Parameter input and through the Overall Parameter input section)
and can be used to smooth the match. However, caution should be exercised because pile damping can affect
the apparent wave speed in the pile. Note that the j p parameter has the same value for all pile segments. It is
preset to 1% for steel piles and 2% for concrete. For timber piles or when a concrete pile exhibit s severe
cracking, a somewhat higher value may be acceptable.

2.7 Non-uniform Piles and Impedance Variations


A pile with some limited damage or other deviation from the theoretical pile profile may be modeled with
impedance variations. Impedance variations can either be entered in CAPWAP’s pile model (PM) input section
by modifying the basic pile profile quantities cross sectional area, elastic modulus and specific weight, or in the
Impedance input section. The latter method more easily facilitates changes to the model for situations that
differ from the designed or expected pile behavior such as bulges or necks of cast -in-place piles.

If the impedance modification is done by modifying the pile profile (either by changing area, elastic modulus or
density; user may also need to insert a line where the change is to be made) in the PM input section, then the
user should also change the perimeter in the PM profile input, if that impedance change indeed affects the
change of piles surface area and volume. Otherwise the unit resistance values would not be correctly
calculated.

If the change is done in the Impedance menu, by Added Impedance, then the user has the choice of
modifying the shaft perimeter and bottom area such that the unit resistance values reflect t he pile’s
proportional surface area change corresponding to the changed impedance values. This can only be done in
the third page of the PM input. The program will in that case also display the corresponding change of pile
volume. The assumption is, therefore, that the impedance change was caused by only a cross sectional area
change and not by an elastic modulus or mass density change. For that reason, there is no change in wave
speed.

Denoting with an * the modified quantities, after changing the impeda nce of th i th segment from Z i to Z*i then
the modified cross sectional area is

Ai * = Ai Zi */Zi (2.27)

The new perimeter is

Pi * = 2(πAi *)½ (2.28)

and the new volume of a segment is

CAPWAP Background Report Version 2014 19


Vi * = Ai * ΔLi (2.29)

Summing up all N p segment volume values (modified or not) yields the total volume of the pile which can be
compared with the installed concrete volume of the pile. The Volume Ratio is defined as the volume computed
from the cross sectional areas of the pile profile, A i , divided by the volume computed from the modified areas
Ai *, i.e.,

Vi * / Vi = Σ(Ai *ΔL i ) / Σ(Ai ΔLi ) (2.30)

Of course, for a preformed pile such as shown in Figure 2.5.1, these surface area and volume changes are of
no interest and should not be used for unit resistance calculation. Please see the Help description of the
segment properties option in the pile model display of the PM option).

Finally it should be noted that the stiffness values used in the static analysis will always be calculated from the
basic pile profile properties.

It is therefore recommended to make the stiffness values consistent with the modified impedance and thus A i *
values, and the changed impedance values have to be used to convert the pile profile to the new values. This
can be done on the second page of PM by clicking on Update Profile Using Added Imp. (on the first page - it
may be necessary to click on the Non-uniform button.)

2.8 Overall Wave Speed


Pile material properties are not always constant over the pile length. For example several horizontal hairline
cracks may exist in a concrete pile and they would slow the wave propagati on, however, the pile top where we
attach the sensors, would have the wave speed of the intact concrete. Thus, to match the delayed pile toe
response without changing the pile top properties (which are linked to the determination of the pile force) it
would be necessary to either insert dispersed tension slacks along the pile (numerically not a very good
model) or reduce the elastic modulus along the pile length which in turn would create reflections due to the
reduced impedance values. A better and much simpler solution would be to reduce the overall wave speed by
increasing the computational time increment, Δt. This option is offered on the first page of the pile model ( PM)
input section of CAPWAP. Note that this is only a reasonable adjustment if it is ce rtain, for example because
of a good force-velocity proportionality, that the wave speed of the pile top material is indeed different from the
overall wave speed. Note that the overall wave speed is not always lower than the pile top wave speed for
concrete piles. Particularly cast in place piles may have a concrete of higher strength, elastic modulus and
wave speed at depth than near the top, because of the higher pressure under which the bottom concrete
cures, or due to a lower strength cast extension at the pile top

The wave speed adjustment should not be used if the length of the pile is uncertain; for example instead of
increasing the overall wave speed it should be checked whether or not pile toe damage existed. Hopefully
early driving records would be available that would indicate the wave speed for the definitely undamaged pile.

If tension cracking causes the wave speed change then the most likely effect is a much lower wave speed
while the pile is in tension than prior to a tension wave developing , i.e. in an easy driving case at a time L/c
after impact. This means that the wave speed in the pile changes at some time during the dynamic event. The
2014 version introduced option, described in Section 2.8, called Variable Time Increment may be more
appropriate to use than the overall wave speed change which modifies the wave speed for all times of the
event.

Another way of producing a modified (but only lower) wave speed is by adding plug mass. The CAPWAP
software allows for both toe plug and shaft plug modeling which in effect causes a soil resistance effect
proportional to acceleration. Because the effect of plug mass is similar to giving the pile a greater mass
density, the wave speed of the pile model with plug mass will be lower than one that has no plug mass (e.g.
the apparent 2L/c will be delayed). The plug resistance is further described in the soil model section of this
report.

2.9 Variable Time Increment


(Actually this option could also be referred to as variable wave speed).

For piles which experience hairline tension cracking during driving or testing, the wave speed may be
noticeably lower when the stress wave becomes tensile. In effect this means that the overall dynamic elastic

CAPWAP Background Report Version 2014 20


modulus is different after the first L/c time when tension develops in the pile. Tension slack modeling is one
way to improve the match, however, for distributed cracks multiple tension slacks would be needed and that
may lead to rather rough computed pile top curves (e.g. instability or large oscillations in the computed curve).
Also, it is a lengthy process and the right combination of segment number, slack size and slack efficiency
values can only be determined in a tedious trial and error procedure. Alternatively, one would want to use a
lower Overall Wave Speed in PM (corresponding to a reduced WC in the PDA software). However, that would
change the wave speed already in the first part of the record when the pile still behaves elastically because it
is fully in compression. Furthermore, once stress levels decrease, the pile may again behave elastically and
then the original wave speed and thus the original time increment would again be the correct choice.

To help in cases where the wave speed changes after the initial, compressive impact event, the user can now
specify a new time increment to be used after a certain time using the Variable Time Increment option.

Please note that the variable time increment affects the way the signals are displayed on the screen. For
example the tick marks indicating the full effect of the soil resistance components. Since this involves some
complex scaling calculations, the screen display may not be current during automatic searches. However, it
will be correct after the search is finished.

Time increment changes of as much as 100% (ratio = 2) have been seen. Note, that this would mean a wave
speed that is one half of the pile’s elastic wave speed. If this were a change that would actually be considered
for the pile material between cracks, the resulting elastic modulus would be ¼ of the elastic value and that
would rarely be the case. However, such a change of reduced apparent wave speed is potentially reasonable
when the opening of several cracks is considered. Note that this wave speed change in general only happens
from time L/c to time 2L/c after impact in the pile and that its effect is only apparent after 2L/c at the pile top
(or sensor location). In other words, it would not be possible to check the wave speed change by timing the
arrival of the upward traveling wave. Figure 2.9.1 shows force, velocity, force in wave-up and displacement for
a concrete pile whose wave return at and after time 2L/c is rather undefined. In fact, the velocity increases
(and therefore the wave-up decreases) at and after 2L/c after impact at a very low rate and rather than
reaching a peak similar to and at time 2L/c after the impact peak, has a very flat top . This is typical for a pile
with tension cracks. (The fact that the wave-up curve decreases already at an early time has to be attributed to
an early unloading effect (the velocity goes negative rather early).

There have been occasions when it appeared as though a time increment increase is also indicated for open
ended steel pipe piles which have a moving plug. For certain times the plug may move with the pile and by
adding mass it would cause a lower natural frequency of the system. In this case the st arting time for the
increased time increment would probably be much later than 2L/c thereby only causing an improved match but
not a significant change of results. Also should the plug move again it may be desirable to turn of f the changed
time increment and, again, for that reason two or more time increments and time increment ratio inputs may be
needed can be entered.

CAPWAP Background Report Version 2014 21


Figure 2.9.1 Example record of a slow wave return which may be
caused by tension cracks

CAPWAP Background Report Version 2014 22


3 THE CAPWAP SOIL MODEL

3.1 Basic Relationships


CAPWAP uses a modified Smith approach to calculate the resistance acting on a pile segment i during the
dynamic event.

The displacement and velocity of each pile segment relative to the soil is the basis for computing the soil
resistance forces (Note: in the Smith approach, the soil surrounding the soil -pile interface is considered fixed;
an extension to that fixed soil assumption is CAPWAP’s radiation damping model which gives the soil one
degree of freedom). The soil model consists of an elasto-plastic spring and a linear dashpot (Figure 3.1.1)
described by three parameters at every soil segment k: ultimate resistance R uk, quake q k, and viscous
damping factor J k. The total static bearing capacity R u of the pile is the sum of all R uk. The total (static plus
dynamic) resistance force R k at soil segment k is computed from

Rk = Rsk + Rdk (3.1)

where Rsk and Rdk are the time varying static and dynamic soil resistance forces at soil segment k which may
act against the motion of pile segment i. The static resistance is defined as being displacement dependent; the
dynamic or damping resistance varies with the velocity. Added together they are the total soil resistance.

Figure 3.1.1 The CAPWAP soil resistance model is based on Smith


with modifications

Soil resistance forces may act at each embedded pile segment. However, since the pile segments are usually
rather short and because of the limited resolution of the analysis approach, it may be sufficient to have one
soil resistance element at the bottom pile segment for end bearing and one shaft resistance element at every
second pile segment for the portion of pile embedded in the soil (this better matches the resolution of the rise
time – time from initial impact to the first peak in velocity). Thus, the number of pile segments N p may be
greater than or equal to the number of shaft resistance elements N s.

The soil resistance R k which acts on the pile segment i can be calculated using following equations (before
rebound occurs):

Rk = k sk u i + J k(vi ) for u i < q k (3.2a)

Rk = Ruk + J k(vi ) for ui ≥ q k (3.2b)

Where:
vi - velocity of the pile segment i;
u i - displacement of the pile segment i;
k sk - the stiffness of the soil in the pile-soil interface at soil segment k;
Ruk - the ultimate static resistance at soil segment k;

CAPWAP Background Report Version 2014 23


q k – the quake at soil segment k, q k = Ruk/k sk;
J k - the viscous damping factor at soil segment k.

3.2 The Static Soil Resistance, Rsk


The static resistance R sk is represented by a linear spring with a slider element (see Figure 3.1.1) which limits
the spring force to the ultimate resistance R uk during loading and to R nk during unloading. Thus,

Rnk ≤ Rsk ≤ Ruk (3.3)

The sum of all R uk values is the ultimate compression shaft resistance of the pile (depending on site conditions
and pile length, a large portion of the ultimate shaft resistance, e.g., 80%, may be available for uplift). Adding
the ultimate toe resistance, R ut, to the ultimate shaft resistance yields the total ultimate capacity of the pile. It
is a requirement in CAPWAP that the predicted ultimate capacity of the pile is fully mobilized during the
dynamic loading event (see also the discussion in Section 3.2.2).

CAPWAP actually allows the user to model not only one but three different toe resistance values. The first toe
resistance is always acting against the bottom of the pile. The second and third toe resistance, if they are
given a non-zero value can be assigned to any of the shaft resistance elements. This is useful for situations
with reductions of pile perimeter and thus intermediate toe resistance (e.g. if an H section protrudes out the
bottom of a solid concrete pile). The second and third toe resistance can also be considered part of the bottom
end bearing model and allow for a non-bilinear modeling of a load-displacement behavior (see Figure 3.2.1 A
and B). The total calculated CAPWAP capacity is, therefore,

N
Ru = ∑k=1
s
Ruk + Rut + Rut2 + Rut3 (3.4)

Figure 3.2.1 Options for multiple toe resistance modeling


(A) combining at the bottom and (B) individual toe
resistance effects at different pile segments

3.2.1 The Unloading Limit, UN


According to Smith, (Figure 3.2.2) for the shaft resistance R nk = -Ruk, however, in CAPWAP the negative shaft
resistance limit is variable.

Rnk = -Un Ruk (3.5)

For the shaft resistance, the factor U n can vary between 0 and 1; it is a CAPWAP input parameter where it is
designated as UN. The same UN value pertains to all shaft resistance segments. While the Smith assumption
of Un being equal to 1 may be reasonable in the static case (tensile or uplift friction equal compressi ve or
downward friction), during dynamic events U n oftentimes has to be reduced to values much less than 1 (see
Figure 3.2.3). The reason may be an imperfectly rigid soil surroundinging the pile or residual stress effects. If
the UN value were not adjustable then the late part of the record, when the pile rebounds, would show very

CAPWAP Background Report Version 2014 24


low or negative upward velocity values. In easy driving situations, when there is no rebound, the quantity UN
has little to no effect. Note that for residual stress analyses, UN should not be very small (say less than 0.2) or
there would be no significant negative resistance forces and therefore no residual stress effects.

For end bearing Rnk is always zero.

3.2.2 Loading and Unloading Quakes, QS, QT, CS, CT


As long as the static resistance is between the positive and negative limits, the static resistance is linearly
related to displacement. The static resistance can therefore be calculated from

Rsk,j = Rsk,j-1 +Δu i k sk (3.6)

Where j and j-1 indicate two consecutive time increments during which the pile segment moves a distance Δu i.
With k sk being the soil stiffness which is equal to

k sk = Ruk / q k (3.7a)

during loading (positive or downward pile segment velocity) and

k sku = Ruk / (q k c u) (3.7b)

during unloading (negative or upward pile segment velocity), where c u is the quake unloading multiplier.

The quantity, k sk, in Eq. 3.7a is the soil stiffness of the k-th resistance for increasing displacements or positive
(downward) velocities. Reversely, the so-called “quake” which is the pile elastic displacement limit after which
the elastic spring goes plastic is given by

q k = Ruk /k sk (3.8)

with q k being the actual loading quake (QS for shaft and QT for toe in the CAPWAP parameter input). Quakes
physically and numerically cannot be zero (as equation 3.8 then becomes undefined). Under extreme
circumstances the toe quake may reach as large as 25 mm (1 inch) and the shaft quake maxima may be
roughly 8 mm (0.3 inches). However, most commonly shaft quakes are in the neighborhood of 2.5 to 5 mm
(0.1 to 0.2 inches). Note that the larger the quakes, the longer it takes for resistance to get activated which
means that the computed upward wave is not increasing as quickly as for smaller quakes; larger quakes also
unload slower.

As mentioned previously, it is a basic requirement of the CAPWAP approach (and dynamic testing in general)
that all static resistance components included in the calculated ultimate capacity, R u, are fully activated. This
requirement can only be met if the displacement of a pile segment i is greater than the quake of the resistance
acting at that segment. For this reason, the recommended upper limit of the shaft or toe quake is equal to the
corresponding pile displacement.

Shaft quakes are usually kept the same for all shaft segments, however, they can be varied. This is a useful
feature when the displacements near the bottom of the pile are rather small requiring quakes on those
segments to be smaller than the quakes in the upper part of the pile for full activation of the predicted
resistance values. (A possible consequence of working with small quakes because of small pile displacements
is that the activated resistance is less than actually available resistance . Better resistance activation would
require that greater energies are delivered to the test pile and that may not always be possible. In any event,
extrapolation of resistance values corresponding to higher quakes might be tempting but is, however,
speculative and cannot be recommended as it is a non-conservative assumption).

For negative (upward, rebound) pile velocities, the stiffness of the soil resistance may be different from the
loading quake as shown by Eq. 3.7b. The difference is the quake unloading multiplier c u which is often less
than 1 making the unloading stiffness greater than the loading stiffness (see Figure 3.2.2 for the shaft
resistance and Figure 3.2.3 for the toe resistance).

The factor c u is represented in CAPWAP by input parameters CS and CT for shaft and toe, respectively. The
multiplier default is 1.0 which makes loading and unloading quakes eq ual. The same unloading multiplier
value CS is applied to all skin quakes. A low unloading quake causes a quick shedding of load during rebound
and therefore lowers the computed upward wave in the late part of the record. To avoid zero unloading
quakes, which are a numerical impossibility, CS and CT cannot be zero, in fact, they rarely should be less than

CAPWAP Background Report Version 2014 25


0.3 and never less than 0.1 unless the loading quake itself is unusually large. Occasionally, the match requires
that either CS or CT is set to values greater than 1.0. This is plausible, for example, when the toe gap (input
parameter TG - see below) is greater than 0; in effect, the quake at the pile toe is then the sum of TG and QT
and thus greater than QT and the unloading quake can then also be greater than QT. Another situation where
it is reasonable that CS and CT are chosen greater than 1 is when the radiation damping model is employed.
The larger CS and CT values are then compensating for the very simplified model of the soil surrounding the
pile-soil interface.

Figure 3.2.2 The basic resistance force vs pile displacement


model by Smith

Figure 3.2.3 The CAPWAP modified shaft resistance vs pile


segment displacement model

3.2.3 The Reloading Level, LS and LT


The reloading option specifies the resistance level during a second or later loading cycle where the loading
quake switches from the unloading level to the reloading level (see Figure 3.2.3). Expressed mathematically for
a shaft resistance R k the reloading quake is given by:

q kr = qkCk for Rsk < L s Ruk (3.9a)

and

q kr = qk for Rsk ≥ L s Ruk (3.9b).

or in words, "Below the reloading level, L s Ruk (for the shaft) or L t Rut (for the toe), the soil stiffness equals the
soil's unloading stiffness in a second or later loading cycle within the same blow." The corresponding input
parameters are LS (LS is the same for all shaft segments) and LT. The reloading levels are normally set to 1.0

CAPWAP Background Report Version 2014 26


which means that reloading is occurring at a stiffness given by the unloading quake. These parameters only
infrequently need adjustment (they can be adjusted in Overall Parameters). Since they usually only affect the
late record portion they have hardly any influence on the bearing capacity prediction. However, LS or LT
values less than the default of 1 may help improve the match when the late record portion shows high
calculated wave up values. In that case one would typically try values of zero for either LS or LT and/or LS = -
1. (LT < 0 does not exist because R st is never negative.) Note that as long as CS (CT) is 1, LS (LT) has no
effect. However, if the unloading quakes are smaller than the loading quakes then reloading levels less than 1
tend to lower the late computed wave up curve, because of a slower reactivation of static resistance.

Figure 3.2.4 The CAPWAP modified toe resistance vs pile


toe displacement model

3.2.4 The Toe Gap, TG


Piles driven to a hard end bearing layer may, during rebound, separate from the end bearing layer a short
distance after the end of the previous hammer blow. The next blow then has to move the pile toe through that
distance before it starts to contact the bearing surface again and activates resistance. Thus under the next
hammer blow the resistance remains near zero until the pile toe has closed this gap and again reached the
bearing layer. To model the zero or very low resistance a toe gap, g t, (TG is the CAPWAP input parameter)
has been added to the static toe resistance model (see also Figure 3.2.4). Note that for full resistance
activation, the sum of the toe gap plus toe quake must be less than the maximum pile toe displacement of the
blow. The static toe resistance, subject to the gap g t, is therefore during initial loading

Rsk = 0 for u i < g t and


Rsk = Ruk for u i > g t = g t (3.10)

where k is equal to N s + 1 (pile toe) and i equals N p (pile bottom segment) R sk is zero for displacements less
than the gap and equal to R uk for displacements greater than the sum of gap and toe quake. During unloading,
the toe resistance follows the unloading quake. In general, for wave equation analyses and therefore also in
the CAPWAP output, the gap is added to the toe quake.

Notes:
1. The soil gap is not a soil property, but rather a randomly occurring pile toe movement condition.
2. All three toes can have individually different toe gaps in the 2014 CAPWAP version.

3.3 Dynamic Soil Resistance or Soil Damping


Viscous resistance forces, considered a function of pile velocity only in the basic soil model, also resist pile
penetration and are responsible for a relatively quick decay of pile motions and decreasing amplitudes of
vibration. In fact comparing typical wave equation analyses with measured records it is frequently found that
the Smith damping approach actually produces less of a damping effect than would be observed in the
records. For that reason, CAPWAP works primarily with a Smith-viscous model, represented by a linear
dashpot (Figure 3.1.1), rather than the tradition Smith approach.

CAPWAP Background Report Version 2014 27


3.3.1 Smith Damping, SS, ST
The traditional Smith wave equation definition for damping at the k th soil segment is

Rdk = J sk vi Rsk (3.11)

which makes the dynamic resistance, R dk, dependent upon both segment velocity, v i , and temporary static
segment resistance, R sk, by a dimensional damping Smith factor J sk (units are s/m or s/ft) for soil element k.
However, as mentioned, for signal matching it is more realistic to use linear Smith -viscous approach:

Rdk = J sk vi Ruk (3.12)

Which uses the same damping factors as the basic Smith approach, but since R uk is a constant, the damping
factor is not diminishing with the static resistance and thus makes for a greater damping effect. In general, it is
sufficient to assign the same damping factor (input parameter SS to the shaft and another one, ST, to the toe.
Varying the damping factors along the shaft is sometimes desirable and can be accomplished by either
damping multipliers or by added damping (see below).

Smith damping factors have the advantage over other formulations (or non-dimensionalizations by pile
impedance as the Case damping factor suggests) in that their magnitudes have relatively modest ranges. In
CAPWAP the recommended range is 0.08 to 1.3 s/m or 0.024 to 0.4 s/ft). However, one has to recognize that
these recommendations are for the usually encountered pile driving velocities and for at most moderately
plastic soil types and typical pile driving velocity ranges. For low velocities (e.g., less than 0.3 m/s or 1 ft/s)
research performed by Coyle and Gibson (1970) suggests that damping forces should be calculated according
with an expression like R d = J CGvn. The exponent n was found to be variable but frequently in the
neighborhood of 0.2. CAPWAP-2014 acknowledges this by increasing the recommended upper limit for SS
and ST from 1.31 to 1.31[(3.28max(vtop )]^-0.8 s/m or from 0.40 to 0.40(max(v top )-0.8 ) s/ft for max(v top ) values
below 0.3 m/s or 1 ft/s. Below max(v top ) = 0.03 m/s or 0.1 ft/s the upper limit recommendations for SS or ST
are not further increased. The distinguishing velocity value max(v top ) is the maximum measured pile top
velocity in m/s or ft/s, respectively.

3.3.2 Case Damping, JS, JT


Traditionally, CAPWAP damping input was made with the so-called Case Damping factors, J c, (input
parameters) JS and JT, which are non-dimensionalized by dividing a purely viscous damping factor (units
kN/m/s or kips/ft/s) by the pile segment impedance which has the same units. It is easier, however, to
recommend a reasonable range of Smith damping than Case Damping factors and for that reason, the Smith -
viscous approach has become the preferred input mode. Conversion from one damping definition to the other
is straight forward:

J c = J sk Ruk / Zi (3.13)

where J sk, Ruk and Z i are, respectively, the Smith damping factor, the static ultimate resistance and the
impedance of one segment (shaft or toe).

In CAPWAP, damping input can be specified by the non-dimensional Case factor J c (JS or JT) or with the
Smith-viscous damping factor, J sk (SS or ST). However, any change of static resistance will modify J c while
J sk remains unaffected.

3.3.3 Damping Options, SO, OP


CAPWAP normally calculates damping forces according to Eq. 3.12. However, the user can also switch to the
standard Smith equation for both shaft and pile toe. Particularly for the pile toe (and large toe quakes) the
original Smith calculation of Eq. 3.11 sometimes yields better signal matches at the time of the wave return.
For that reason, CAPWAP offers the following 3 damping options which can be selected with options SO and
OP for shaft and toe, respectively (the SO option however is normally used with the standard Smith viscous
approach):

SO = 0, OP = 0:
This is the standard Smith-viscous option (Eq. 3.12) and is most commonly used for the shaft.
SO = 1, OP = 1:
This is the traditional Smith option (Eq. 3.11), which in CAPWAP is rarely used,
SO = 2, OP = 2:
This approach uses Eq. 3.11 until full activation of the static resistance has been reached and Eq.
3.12 for all times after that. OP = 2 is frequently employed for modeling toe damping and is
particularly useful when the Toe Gap option has to be used or for larger quakes. OP = 2 has been

CAPWAP Background Report Version 2014 28


made the default in 2014. However, the user is urged to try to improve the match by also working with
OP = 0.

3.3.4 Plugs (Acceleration Dependent Resistance), PS, PL


While not velocity dependent, this resistance component is also a dynamic resistance. It assume s that
additional mass(es) is (are) either attached to certain pile segments or to the toe. A toe plug can also be
commonly used to model the end plate for a closed end pipe pile. For example, a mass of the soil could be
sticking to the pile or it may be trapped underneath the bottom of a displacement pile (this is probably rather
common). For open ended pipe piles or for H-piles, a soil plug may develop inside the pipe or between the
flanges. Such a soil mass, if it is moving with the pile, develops an inertia force which in turn causes a
compressive reflection.

Whether located at the pile toe or somewhere along the shaft, this inertia-type resistance delays a tension
wave reflection form the toe and, therefore, makes the pile appear longer than it is (the toe reflection occurs
later if such a mass exists). In effect, the mass increases the ρ in the equation for the wave speed, c = (E /
ρ)0.5 . Thus, it is conceivable that the soil resistance effect of added soil mass could also be modeled by an
impedance increase plus either a pile length increase or a wave speed decrease. (Of course, a pile impedance
increase by itself, however, does not cause any change in wave speed.)

The soil mass, m p, (PS and PL, with units of weight) are the CAPWAP input parameters) exerts an external
resistance force (R mi ), acting against a pile segment, i, or the pile bottom (see Figure 3.3.1).

R mi = a i m p (3.14)

where a i is the acceleration of the pile segment i where the inertia resistance acts.

To prevent that the inertia resistance smoothens out the wave too much and then appears like a damping
resistance, PS and PL should be limited to at most 3 times the weight of the pile segment to which they are
attached. Most commonly PS and PL remain at zero.

The plug masses can be considered a pile model variation although they represent a soil weight. For that
reason N p PS values can be modeled. The PS value displayed among the CAPWAP input parameters is the
non-zero plug weight located closest to the pile top; changing this value will modify all other PS values
proportionally. Input is most likely, however, done in the drop down menu Plug on Shaft. Note that in contrast
to other soil resistance input parameters, PS can be individually specified for all pile segments.

3.3.5 Radiation Damping, SK, BT


Smith type soil resistance only considers the pile motion when calculating resistance forces. Of course, the
resistance forces in the pile-soil interface exert a reaction force on the soil which causes the soil surrounding
the pile to deform and therefore displace. Soil movements have the greatest effect on the soil resistance
forces when the pile motions are small such that a true shear failure does not occur. An example is a drilled
shaft which is socketed into hard rock. As the pile exerts compression forces against the rock, a wave is
generated in the rock (which causes particle motions, i.e., velocities and displacements of the rock) and the
soil resistance appears to be a function of velocity rather than displacement. This example explains why we
talk about "Radiation Damping" (the energy is radiated away from the pile rather th an consumed for soil
shearing) and why we use a mass and a dashpot to replace the rigid soil support of the Smith model.

Another example where the soil motion happens, and if not recognized could give poor resistance results, is a
pile with a high wave-up and thus a high apparent friction during the first 2L/c time period (i.e. before the
impact wave returns to the pile top after reflection at the pile bottom). However, a very low wave -up curve after
the 2L/c wave-up valley suggests a low capacity. Apparently, the static soil resistance like the damping
resistance, decayed quickly, possibly and probably because the soil moved away from the resistance as soon
as the rather high resistance was activated. Based on the standard Smith model, CAPWAP would normall y
predict low static and unusually high damping resistance as evidenced by large Smith damping factors (say, in
excess of 0.4 s/ft or 1.3 s/m). A radiation damping model to represent soil motion can be used in such cases to
obtain reasonable comparisons with static load test results by limiting the maximum Smith damping factor for
the shaft to 0.4 s/ft (1.3 s/m) (Likins et al. 1996).

Radiation damping is also frequently employed for the analysis of open ended pipe piles driven into granular
soils with partial plugging effects (the plug at first moves with the pile during a blow but then begins to move).
Thus internal friction forces at first are high, but diminish quickly, independent of the pile displacement. If this
phenomenon is not modeled with radiation damping, it could lead to underpredictions. However, caution must
be exercised in the case of cohesive soils near the bottom where radiation damping could lead to
overpredictions of end bearing (see also Chapter 7).

CAPWAP Background Report Version 2014 29


Figure 3.3.1 shows the radiation damping mechanism below the Smith mass-dashpot-spring resistance model
for both shaft and toe. Of course, the pile may not support any static load if the soil support dashpot has a
damping factor less than infinity. This philosophical dilemma may be resolved by assuming that J SK or J BT the
damping factors of the soil support, are less than infinity only during the dynamic event.

The shaft soil support masses, m RD, and damping factors J RD shown in Figure 3.3.1, appear to establish a good
model for energy dissipating waves radiating into the soil surrounding the soil -pile interface. The governing
equations only change in that the pile segment velocity, v, and displacement, u, are repl aced by the relative
quantities, vr and u r . The velocity v ss and displacement u ss of the i-th radiation damping mass at time j are
calculated from

vss,i,j = vss,i,j-1 + (Rk + J vk vi – J RD vss,i,j-1) / (J RD + m RD/Δt) (3.15)


and

u ss,i,j = uss,i,j-1 + (vss,i,j-1) Δt (3.16)

Figure 3.3.1 The complete CAPWAP soil model for the toe and for two
shaft resistance segments near the pile toe.

where "ss" stands for the motion variable of a soil support mass at any pile segment. For the toe R t and J t can
take the place of R k and J vk which are the Smith static soil resistance and the viscous damping parameter for
soil segment k at pile segment i, respectively. Omitting the time counter j, the relative motions be tween pile
and soil mass can now be calculated from

u ri = ui - u ss,i (3.17)

and

vri = vi - vss,i (3.18)

For convenience, radiation damper values, J RD, are input in CAPWAP as SK and BT for shaft and toe,
respectively, in pile EA/c units, i.e. they are dimensionless. For the radiation damping mass or soil support
mass values, m RD, input parameters MS and MT, default values are calculated by CAPWAP based on the pile
perimeter considerations described and correlation work performed by Likins (1996.The default values
correspond to the weight of a cylinder of soil surrounding the pile and being of roughly 0.3 m (1 ft) thickness.
While the user can enter different values in Overall Parameters these values must not be chosen too small or
unreasonable results could be obtained. Similarly very small J RD values must be avoided while making them
larger than the upper bound help recommendation is generally of no particular concern. Note that only one and
therefore the same SK and MS values can be entered for all shaft segments.

CAPWAP Background Report Version 2014 30


4 OTHER CAPWAP CALCULATIONS

4.1 Match Quantity and Reflections at the Pile Top


At the pile top, either force, F m,j, velocity v m,j, or wave down F dm, j are prescribed (“m” for "measured") at time j.
Then the complementary computed (“c” for "computed") quantity is either the force

Fc,j=Z1v m,j+2F uc1,j (4.1)

or the velocity

vc,j = [F m,j - 2Fuc1,j]/Z1 (4.2)

or the wave up in the first pile segment F uc1, j.

Fuc1, j is readily known from the algorithm calculating the forces in all upwards and downwards traveling wave
values at all times (Chapter 2). The impedance of the top segment, Z 1, is also known from Chapter 2. Note that
the pile top impedance value must have the same value as entered in the PDA, because it is composed of the
area and modulus values on which the strain to pile top force conversion has been based. For that reason, a
modification of the pile top impedance value can only be done in the PDA program.

Obviously, either force matching (comparing F c,j with F m,j), velocity matching (comparing v c,j with v m,j) or wave
matching (comparing Fuc1, j with Fum1, j = ½[F m,j - v m,j Z1 ]) is possible and can be selected by the CAPWAP
analyst.

Wave matching has the advantage that phase shifts, due to an inaccurate wave speed assumption, are easily
detected and avoided. The wave-up also contains only the upward traveling waves due to the soil resistance
or any pile non-uniformities (including the pile toe) and does not contain any downward wave components
such as the hammer imposed impact wave or top reflected downward waves. Wave matching greatly simplifies
the visualization and allows for a better resolution of the soil resistance effects. Wave-up matching is,
therefore, the preferred analysis type.

One word of caution: if one of the two measured quantities has a significant flaw then it should not be used for
the pile top boundary input condition, or not even as a part of the measured wave -down. For example, let us
assume that the measured force curve contains a spike due to an electronic problem, but appears otherwise to
be a satisfactory record. Velocity matching would impose the questionable force record as a top boundary
input condition and the spike would cause a false disturbance. Even the wave-down record would contain one
half of the spike and affect the calculation. However, by using the velocity as the top boundary input condition
and matching the force, this would not cause a problem in the calculation and the mismatch at the spike could
be simply ignored. In other words, it is wise to use as an analysis input only a quality record component and
match the less perfect record component, ignoring mismatches at record defects. Ideally, ho wever, a better
record (an impact without a flaw) is selected for the analysis.

While final results generally show the calculated and measured match quantities, for wave matching, the
"calculated" top force is often compared with the measured top force. However, the computed top force now
has to be determined (like all other forces in the pile) by superposition of the downward and upward wave in
the upper segment. Thus, for wave matching one compares the measured force F m,j with

Fc,j = Fdm,j + Fuc1, j (4.3a)

which can be rewritten as

Fc,j = ½ (F m,j + Z1 v m,j) + F uc1, j (4.3b)

Thus, the standard CAPWAP match presentation, based on wave matching, compares a computed force,
which includes 50% of the measured force, with the measured force. Obviously wave matching therefore
produces visually better looking matches than either velocity or force matching.

4.2 Pile Toe Reflections


At the pile toe, force equilibrium requires that

FuNp,j = -FdNp,j-1 + RNs + RNs+1 (4.4)

CAPWAP Background Report Version 2014 31


with R Ns and R Ns+1 denoting the shaft resistance at the last soil resistance segment and the toe resistance
(resistances include static plus dynamic plus plug inertia), respectively.

4.3 The Residual Stress Analysis (RSA)


The background of Residual Stress Analysis theory is described in the GRLWEAP manual and is therefore
considered necessary background reading, and is therefore not repeated. Here no further explanation is given
for the occurrence of residual stresses in pile and soil at the end of a hammer blow except to say that the pile
compresses during impact and then during rebound the soil resists the pile's full extension, thereby
maintaining a compressive force in the pile shaft even after the hammer blow. Longer flexible piles with well
distributed shaft resistance and modest end bearing percentages are more susceptible to RSA than short, stiff
piles or piles with little shaft resistance (piles with no shaft resistance cannot have residual stresses). It was
observed in earlier wave equation studies that residual stresses produce lowe r calculated blow counts than
those from conventional analyses since energy stored in the pile at the end of a blow is available to do useful
work during the next blow.

In the past, CAPWAP engineers sometimes felt that the calculated resistance distributi ons may have been
distorted (compared with their expectations from static analysis for example) because residual stresses were
not considered. The reason is that at the end of a blow some of the upper soil resistance forces are directed
downwards (negative), while portions of the lower friction and the end bearing are still directed upwards.
During the next blow, activation of the upper resistance forces requires pile displacements which first bring the
negative friction forces to zero before positive resistance is generated, making the upper resistance appear
larger. At and near the pile toe, resistance which has been partially preloaded by the previous blow takes less
pile displacement to activate the full ultimate resistance of the lower soil strata, ma king the lower resistance
appear smaller. Since the conventional CAPWAP analysis assumes resistances and displacements initially to
be zero, it may underpredict the lower shaft resistance and end bearing resistances and overpredict resistance
in the upper strata. By performing residual stress analyses in CAPWAP, estimation of the resistance
distribution can be improved, although the predicted total capacity is generally unchanged.

RSA is initiated when the RSA option is given a value greater than zero. Fo r example, if RSA is set to 4, the
normal CAPWAP trial analysis is followed by four additional analyses which use the stresses in pile and soil
from the end of the previous analyses as initial values. In CAPWAP the following steps are undertaken during
a single trial analysis with the RSA set to a number greater than 1.

• In the initial analysis all pile and soil variables are initialized to zero.

• In all subsequent RSA, velocities and, therefore, damping forces are initialized to zero; displacements
and, therefore, pile forces/stresses and static soil resistance forces are not initialized.

• After each analysis (initial and RSA), CAPWAP performs a static analysis which produces an
equilibrium set of soil resistance forces and the resulting residual pile segm ent displacements.
Different from GRLWEAP, which solves a set of simultaneous equations to obtain the static set of soil
resistance forces and associated pile forces, CAPWAP performs a step-by-step static analysis without
velocities and masses until the final top set equals the measured one. This step-by-step procedure is
similar to that described in Section 4.9.

• CAPWAP calculates the blow count from the final pile top set occurring during the last analysis. Before
each RSA analysis, displacements of all pile segments are reduced by the final set of the pile top
segment (they are in general nonzero). Even though the top segment’s final displacement is , by
definition, equal to the measured set, the static analysis performed after the dynamic one may produce
a different final set and therefore a blow count error. Resistance, quake and unloading parameter
adjustments may be used to reduce that error. Note that all pile segments will have the same final set
if convergence has been achieved.

• CAPWAP displays (click View/Display Variables and move to bottom of table) the pile top sets
(DTopNow, DTopOld) of the last two analyses and the corresponding compression values
(COmpNow, COmpOld). The compression values are the difference between the final top and final
toe displacements. If the compression of the pile remains the same (or, say, within 10%) between two
consecutive analyses, convergence has been achieved. If the percentage change of pile compression
between the last two analyses is high, then convergence has not been achieved and it may be
necessary to perform additional consecutive analyses (increase RSA option). For a successful RSA
convergence, it is even more important than for standard CAPWAP analyses that the analysis is

CAPWAP Background Report Version 2014 32


carried out to a time at which the pile velocities have become zero and the displacement has achieved
the final set value. This TEnd value can be set or changed in the CW option of the PDA software
and/or can be checked or changed in INPUT/BLOW RELATED ADJUSTMENT.

• No residual forces exist when negative shaft resistance is prevented by UN=0 (the shaft resistance
unloading level). Therefore, the UN value must be greater than zero for a meaningful RSA. Although
no correlation studies exist, it is reasonable to use UN ≥ 0.2.

• RSA assumes that repetitive blows produce identical force and velocity records and activate identical
soil resistance forces. This situation typically occurs during driving. During the first couple early
restriking impacts when both impact energies and activated shaft resistance values change, MBA (see
next section) would be the more appropriate analysis method, or wait until at least the third impact for
performing RSA.

Different from GRLWEAP which controls the number of analyses for convergence of RSA, when analyzing with
CAPWAP the user has the responsibility to check that convergence occurs. The output allowing for
convergence confirmation only suggest that users investigate convergence and increase the number of
analyses until pile compression values of successive analyses agree. Under complex circumstances (e.g.
slacks) convergence may not be achievable and then RSA should not be used.

4.4 The Multiple Blow Analysis (MBA)


It was mentioned that RSA is not a reasonable analysis option for the very early restriking impacts. For
example, during early restriking, energies increase and shaft resistance decreases while more end bearing is
activated from blow to blow. Thus, the records of force and velocity vary greatly from blow to blow and at the
same time residual forces build up in the pile. For such a situation MBA would be applicable. CAPWAP can
analyze up to ten different (usually consecutive) records with identical quakes and damping values and with
the same static resistance distribution. Furthermore, to simulate the loss of shaft resistance from blow to blow,
MBA offers the multiplication of the shaft resistance forces with a Resistance Reduction Factor.

Please Note: While the MBA approach is plausible, actually achieving good matches and obtaini ng reasonable
results is quite involved and sometimes not satisfactory. It is often easier to perform individual analyses for
several restrike blows, maybe with RSA, and then combining the results. The envelope of all resistance
distributions thus obtained may provide a more consistent and reliable result than MBA.

The process starts with the first record assuming a zero stress state. After the first record is analyzed, a static
analysis determines the static equilibrium condition of the residual stresses . The dynamic and then static
analysis of the second record and then all remaining records follow in a similar manner.

For each record analyzed, measured and computed pile top quantities are compared to calculate a match
quality number. In addition, a realistic comparison of measured and computed pile top penetration can be
made.

It is assumed that a long term capacity, R uk, is present at pile segment i in the beginning of the Multiple Blow
Analysis. A Capacity Reduction Factor, f rk, (0 ≤ frk ≤ 1) is defined as an indicator of the lowest degree to which
the capacity R uk can degrade. The degraded ultimate shaft resistance, R duk, of soil segment k is calculated
from the long term capacity, R uk, and the cumulative pile segment displacement, u ic, as follows.

Rduk = Ruk for u ic ≤ q k (4.5a)

Rduk = ftk Rduk for q k < uic ≤ κq k (4.5b)

Rduk = frk Rduk for κq k < uic (4.5c)

The temporary capacity reduction factor f tk (frk ≤ f tk ≤ 1), is computed from the accumulated pile set u ic,
decreasing linearly from 1.0 to f rk. The full reduction occurs after the accumulated pile set has reached and
exceeded the Reduction Displacement κq k (κ is an input for MBA; Red Displ is shown on the CAPWAP Input
Screen under Capacity Reduction Factor input form when MBA is activated).

Obviously, the assumption that the capacity degrades fully within several quake displacements is rather
arbitrary. In fact, full degradation may actually not be reached if only a few hammer blows are analyzed.

CAPWAP Background Report Version 2014 33


Notes:
• In the above definition of the degraded ultimate resistance, the cumulative displacement u ic has been
used. Since pile variables are not reset to zero between blows analyzed (differ ent from RSA) the
accumulated displacement is the actual pile penetration achieved by the sum of all blows analyzed.

• For a meaningful MBA, records must be available which represent either energy and/or resistance
variation from blow to blow, as often occurs for the initial blows of a restrike. A constant
energy/constant resistance situation may be analyzed by the traditional CAPWAP or if concerns
regarding locked-in stresses exist by RSA (see Section 4.3).

• Each analysis within the multiple record group is considered a trial as in standard CAPWAP. However,
while the conventional CAPWAP considers each record separately, allowing for different re sults from
blow to blow, MBA adds considerably on the "known" side (records) while adding as unknowns the
resistance reduction factors and the reduction displacement. On the other hand, the match quality is
generally not as good for the assembly of all records analyzed as for all records analyzed individually
because damping and quake values are not individually adjusted. However, rather than matching each
detail of an individual record, the more essential overall record behavior is approximated.

• It is important to realize that MBA is potentially non-conservative since it includes resistance values that
can be degraded with low dynamic energies. These degraded capacity values may also be lost
(degraded) and therefore not be useful during static load applications. Indeed MBA replaces the method
of superposition (see Section 4.9) which may also be non-conservative.

4.5 Blow Count Calculation


In contrast to the wave equation approach, blow count calculations are not essential for a successful CAPWAP
analysis. However, there is evidence (and it is reasonable to suggest) that a good CAPWAP result should also
match the observed blow count (or, inversely, the observed set per blow) reasonably well. Since the observed
blow count is rarely accurate for a particular blow, particularly for the first blows of a restrike, it is fortunate that
the accuracy of the CAPWAP prediction does not rely on the accuracy of the blow count which is one reason
why CAPWAP is much more reliable than wave equation analysis for capacity determination.

CAPWAP always calculates and displays for each trial analysis the blow count, i.e. the inverse of the final set.
The final set is calculated in two ways.

The first method (and this is the traditional one like in the wave equation approach) calculates final set as the
maximum toe displacement minus the weighted average quake.

set 1 = utoe,max - ∑(q k Rk)/RU (4.6)

In this equation, the sum is taken over all soil segments k including the toe. This set calculation is only done
when no blow count matching is employed and the associated blow count, i.e. the inverse of set 1, is then only
displayed for information. It is relatively independent of the acceleration measurement, however, it is very
sensitive to small changes in quake. For that reason it was found necessary to calculate the final set as the
average of the final displacement values in all segments i from 1 to N P, if “blow matching” is employed. Thus,

set 2 = ∑ ui, final / NP (4.7)

To activate this method, check the “Match” box in the upper left of the CAPWAP screen. For blow count
matching (see also next section dealing with the calculation of the match quality number) the set 2 value is
used and also inverted and displayed as blow count. In order to avoid a blow count penalty in the MQ
calculation it is necessary to make acceleration adjustments to the measured top velocity that produce the
final set at the pile top for an extended time period (horizontal tangent). This also means that the record
should be sufficiently long and analyzed until the displacement at the pile top has reached the final value.
Clearly, if there are no residual stresses in the pile (e.g. UN = 0) then all final displacements are equal to the
final pile top set and then the blow count matches the observed one perfectly if the pile top record at the end
of the analysis (check “Blow related adjustment” under the Input drop down menu for length of record
analyzed) integrates out to the correct final set. If the record is too short, then dynamic waves still travel
through the pile making the calculated final set either too low or too high, so this should be avoided. If residual
stresses are present, then the pile is elastically shortened at the end of the record and the average final set
will be less than the pile top set. In this case the calculated blow count will be higher than the blow count value
on which the final pile top set adjustment has been based. Since the final set of the measured top record has

CAPWAP Background Report Version 2014 34


been made equal to the observed set (inverse of blow count) a perfect blow count match is then impossible.
However, set 2 will be generally closer to the top value than set 1.

4.6 Match Quality Assessment


In order to produce an objective measure of the quality of a match, and particularly to assist the automatic
search functions, it is necessary to calculate a match quality number which is independent of a personal or
visual assessment. Also, the match quality should be at an optimum for the final CAPWAP solution. The
CAPWAP MQ number is effectively a “mismatch” indicator: the higher its value, the worse the match.
Therefore, the best match would have the lowest possible number. MQ is calculated in t he following manner.

1. The record is divided into 4 different time periods I, II, III, and IV (see Figure 4.6.1)
I. From the onset of impact, say t o, to 2L/c later, i.e. to to + 2L/c (resistance distribution),
II. From t o + 2L/c to t r + 3 ms later (toe resistance development), where t r is time from t o to velocity peak
(the rise time),
III. From t o + 2L/c to t r + 5 ms later (capacity development),
IV. From t o + 2L/c to 25 ms later (unloading behavior).
Obviously time periods II, III and IV have some overlap and the period near and just following 2L/c will
have added emphasis.

Period I II III
2L/c tr+3 tr+5ms
ms

IV
tr 25ms

Figure 4.6.1 Match quality time periods with a wave-up option

2. For each of the four time periods, an “error” is calculated as the sum of the absolute values of the
differences between the calculated and measured quantity at each calculated point in time. For example,
for wave-up matching, the absolute value of the sum of the differences between wave -up calculated and
wave up measured is calculated.

3. The four “error” values are added and then averaged by dividing by the sum of the number of points
involved including the overlaps between time periods.

4. The averaged error value is divided by the maximum pile top force (for velocity matching by the pile top
force divided by the pile top impedance) and then multiplied by 100 to make it a percentage. Thus the MQ
is sort of an “average error” in percent of the maximum force.

5. The blow count penalty is added.

The following equation reflects the above procedure.


N
∑4I=1(∑i=1
I |F -F |)
Mi Ci 100
MQ= ( ∑4I=1 NI
) + MQB (4.8)
Fmax

where - number of data points for each time period, I = I, II, III and IV;

CAPWAP Background Report Version 2014 35


MQB - the Blow Count Penalty (if this option is chosen) is calculated as follows.

MQB = |set 2 - set obs| – 1 mm (4.9)

with a lower limit of zero and an upper limit of 0.1|set 2 - set obs|. The “set” values for MQ B are in mm. For set 2
see Eq. 4.7 while set obs is obtained from field observation (inverse of blow count).

In summary, the Match Quality MQ is calculated as the absolute value of the difference between measured
and computed pile top match quantity at all points in time calculated in the 4 periods plus the absolute value of
the absolute values of the difference between observed final set and the calculated final set reduced by 1mm.

It is somewhat unusual that the absolute value of the differences between measured and computed pile top
quantities has been chosen to calculate MQ rather than the square of the differences. The reason is that the
sum of the absolute values are not as sensitive to small phase shifts which would make small wave speed
inaccuracies (e.g., on both sides of the 2L/c valley an excessively large error contribution.

It is impractical to require that MQ is less than a certain absolute limit since each data set has it s own optimum
match. For example, analysis of a pile with very high relative end bearing will not give as low an MQ as a pile
with high shaft resistance. This variability stems from the fact that end bearing conforms less perfectly to the
elasto-plastic model than shaft resistance. Another reason for relatively high MQ values would be a difficult
pile model such as a non-uniform pile (maybe a drilled shaft with unknown variable impedance values). On the
other hand, where the force record includes a large force peak, F max, relative to the wave-up magnitude, the
match quality could be deceivingly good (see Eq. 4.8). Also very long piles with little shaft resistance tend to
make the “error” in Period I rather low and, therefore, CAPWAP actually normalizes the “ error” in that time
period relative to the time 2L/c. Obviously, if the data itself is less than perfect, e.g. because of high bending,
then certain mismatches are already built into the data set and the best MQ will be relatively higher. In
general, however, for most analyses MQ should be less than 4, MQ values above 6 should give reason for
concern and MQ values above 8 are generally unacceptable.

4.7 Interpolation
The Pile Driving Analyzer collects digital data at certain fixed sampling rates such as 5000, 10,000 or 20,000
samples per second (sps), corresponding to sampling time increments of 0.2, 0.1 and 0.05 ms, respectively.
However, a 1 m long steel pile segment requires a computational time increment of somewhat less than 0.195
ms while a concrete segment would have a wave travel time of somewhere between 0.333 and 0.222 ms. It is
therefore necessary to find the force and velocity values at times that correspond to integral multipliers of the
computational time increments. CAPWAP does this by first filtering (time averaging) the measured data with a
filter frequency that is roughly one half of the computational frequency (inverse of the computational time
increment) and then by linear interpolation between the two measured values that bracket the computat ional
values.

It should be noted that the filtering process may reduce some sharp peak values. However, there is not much
gained by making the time increments (and therefore pile length increments) much shorter than the given
sampling frequency. On the other hand, particularly for the analysis of piles with strong non-uniformities,
smaller length segments and thus smaller time increments may help achieve better matches.

4.8 PEBWAP
PEBWAP, the Pile End Bearing Wave Analysis Program, is based on a simple concept: if the soil resistance is
concentrated at the pile toe, then the Case Method resistance is the correct toe resistance solution. Thus, for a
uniform pile, the total resistance at time increment j can be calculated from the measured downward wave (at
a time 1 which is L/c , i.e., at N p points in time, earlier than j, i.e. at time increment j - Np ) and the measured
upward wave (at a time 2 which is L/c later, i.e. at time increment j + N p), using the Case Method formula

RTj = Fd1, j-Np + Fu1, j+Np (4.10)

The toe velocity is then

vtoe,j = {2 F d1,j-Np - RTj } / Z (4.11)

Multiplying the toe velocity by the Case Damping Factor, J c, yields the damping resistance.

RD j = J c {2 F d1,j-Np - RTj } (4.12)

CAPWAP Background Report Version 2014 36


and subtracting RD j from RT j gives us the static resistance, RS j, assuming that the correct Case damping
factor J c had been picked. Thus we get the well known Case Method formula for the static resistance which, in
the case of a pure end bearing pile, is equal to the static toe resistance.

RS j = (1 - J c)Fd1,j-Np + (1 + J c)Fu1,j+Np = Rtoe j (4.13)

Finally, since we know the velocity at the toe, we can readily calculate the toe displacement by straight forward
integration:

u toe,j = utoe,i-1 + vtoe Δt (4.14)

Plotting R toe,j vs u toe,j, point by point, leads to the pile toe resistance vs displacement curve. Again, if all
resistance acts at the pile toe, then the associated pile top displacement is

u top,j = utoe,j +(L/EA) R toej (4.15)

where L, E and A are the length below sensors, cross sectional area and elastic modulus of the uniform pile.
We can therefore plot the pile top load set curve in closed form and this is the PEBWAP solution.

In summary, PEBWAP calculates the total resistance (J c = 0), the toe velocity, the static resistance (for an
assumed J c >0) and the toe displacement, all for the toe as a function of time. The total resistance and the
static resistance values are then all plotted against the pile toe displacement or optionall y against the top
displacement. Like the Case Method, PEBWAP is only applicable to uniform piles.

If we don’t know it, how should we determine the assumed damping factor? For piles bearing in a soil layer,
several damping factors may be tried until a straight line static resistance vs displacement behavior has been
found. This method would correspond to finding a good match with the elasto-plastic soil behavior assumed for
CAPWAP matching. In the example of Figure 4.8.1 this worked out very well for a J c of 0.3. Please note,
however, that not all PDA records yield such a nearly perfect elasto-plastic toe resistance behavior.

Figure 4.8.1 Example of a PEBWAP result

If the proper damping factor has been found, then the static resistance plot shows in an approximate manner a
reasonable quake as that displacement where the elastic line turns plastic (appro ximately 8.5 mm in the above
example). The graph also indicates a total shaft resistance by the resistance force at zero toe displacement
(practically zero in the present example). (Of course, a significant shaft resistance would violate the basic
PEBWAP assumption.) By the way, the RAU capacity appears in this plot as the capacity at the maximum toe
displacement. It is independent of the choice of J c because at that point the toe velocity and therefore the
damping resistance force is zero.

Unfortunately, the PEBWAP analysis is not very often applicable (a) because it requires a uniform pile, (b)

CAPWAP Background Report Version 2014 37


because the examples with low shaft resistance are not very common and (c) because in many instances the
toe resistance does not behave elasto-plastically (for example in hard rock).

4.9 Static Analysis


The static results of the CAPWAP signal matching process are primarily ultimate resistance values and quakes
(and therefore soil stiffness). Together with the pile stiffness, this information is sufficient for the calc ulation of
a load-set curve. CAPWAP produces a simulated pile top load set curve based on such a, sometimes called, t -
z analysis which is then included in the CAPWAP final result page. It is calculated based on the elastic pile
properties E, A, L and the calculated static soil resistance properties R uk and q k. The toe quake includes the
toe gap considering this a peculiar dynamic resistance effect which is generally not present during static
loadings (the option exists to calculate the load set curve with the gap separate from the toe quake).

4.9.1 Analysis
The calculation is based on an assumed toe displacement which is applied in small increments to the pile toe.
The mechanical pile model consists of elastic springs, with stiffness k i , connected together at node points. The
soil is represented by elasto-plastic springs attached at the node points of the pile springs. The applied toe
displacement can be equal to the final toe displacement steps calculated by the final dynamic trial analysis of
CAPWAP or it is an assumed monotonically increasing value (standard). In the latter case, the analysis is
finished when all segments have reached or exceeded the quake displacement. At that point a set equal to the
observed set is added to the pile top load-set curve and a mirror image of the loading curve is added to the
pile top plot to show a reasonable unloading curve. The basic equations for the pile forces, F i , and the pile
displacements ui are simply:

Fi = Fi+1 + Rk (4.16)

u i = ui+1 +Fi /ki (4.17)

with Rk being the static resistance acting at pile segment i. It is calculated in the same way in which it is
calculate in the dynamic case. For the bottom segment N p the calculation is started with u Np prescribed as
mentioned above and F Np being the sum of all toe resistance and shaft resistance forces acting at the bottom
pile segment.

4.9.2 Limitations
It is somewhat daring to generate a “static” load-displacement curve from a dynamic test and it is important
that the limitations of this approach be clearly understood by the analysts as well as their clients.

1. The calculated load-displacement curve represents a load test that was performed during a fraction of a
second. For that reason the result cannot include any displacements which are caused by consolidation of
soil layers below or around the pile. Also any creep settlements would not be included.

2. The calculated load-displacement curve represents the SRD, i.e. static resistance to driving at the time of
testing. Static resistance results from restrike records, obtained after sufficiently long waiting times after
pile driving, generally represent an SRD that corresponds to the long term static capacity. However, while
velocity dependent components have been accounted for by signal matching based on the Smith soil
model, soil sensitivity, large displacements and other effects may cause differences between SRD and long
term static capacity. Additional options allow for consideration of such effects in the “manual” static
analysis.

3. For concrete piles the elastic modulus used in the calculation of the static load-displacement curve is the
same as that determined by the PDA. Static elastic moduli generally are lower than the dynamically
determined ones. The static analysis options allow for input of a pile modul us reduction factor.

4. Often, but not always, static quakes are larger than dynamic ones. Again the static analysis options allow
for a correction by offering an added quake input.

5. Records obtained with relatively low energies and therefore associated small sets (high blow counts)
normally result in a mobilized capacity that is less than the ultimate pile capacity. The calculated load -
displacement usually reflects this fact in a narrow cycle with little or no permanent set.

6. There are many other differences between static and dynamic test conditions which can cause differences

CAPWAP Background Report Version 2014 38


between static and dynamic load-displacement curves. They are too numerous to include in this description
of CAPWAP’s static load-displacement calculation. For example, excavations or scour around the pile
reduce both effective stress and shaft resistance contact area. Variable water table or artesian water
pressures also affect the effective stresses and, therefore, the bearing capacity. Also negative skin friction
effects, due to settlements occurring around the upper pile portion, lead to additional loads and reduced
shaft resistance. These latter effects can be estimated from the CAPWAP calculated resistance
distribution, however, the analyst has to be knowledgeable in geotechnical engineering in general and in
estimating soil settlements in particular. Chapter 7 further discusses limitations of the predicted capacity
and therefore the load-displacement curves.

Please note: In the standard CAPWAP output graphics, the pile top force and the pile bottom force are both
plotted against their respective displacements, i.e., the pile toe force is plotted against the pile toe
displacement. This is different from presentations of actual static load tests where the force at the pile bottom
has been measured (but not the toe displacement) and where the bottom force is then plotted against the pile
top displacement.

The following options for result presentation have been provided for added realism.

4.9.3 Static Analysis Options


A great number of static analysis options make this portion of the CAPWAP program very flexible and allow for
calculating load-displacement curves that are based on calculation methods other than CAPWAP. This would
be done in Output/Static Analysis /Manual Input It is the analysts’ responsibility to use these options
properly and with best judgment.

(a) Analysis Options (Dynamic applied toe displacement option) The standard “Monotonic(ally) Changing Toe
Displacement, E(lasto)-P(lastic) Res(istance) mentioned above comes with several sub-options.
i. Use User Defined Capacity: Users can add a second load-displacement curve with their own
specified capacity. The CAPWAP calculated resistance forces will then be proportionally modified.
ii. Users can Use an Extrapolated Soil Model which requires specifying additional Extrapolation
Options with parameters describing either a strain hardening at the toe or a strain so ftening along
the shaft for displacements which are in excess of the quake values determined during signal
matching.
 The first extrapolation option allows for an increased toe resistance with large settlements under
static loads. This option is thought to show a more realistic load-displacement prediction for
the case of displacement piles where the toe is in coarse grained soils and low dynamic pile
penetrations (high blow counts). The required input value is the ratio of the extrapolated to
CAPWAP calculated end bearing. After quake is reached, an exponential curve is used to
interpolate between the CAPWAP calculated capacity and the extrapolated capacity which is
only reached at an infinitely large displacement (see Figure 4.9.1a). Obviously, the increased
toe resistance option can lead to a serious over prediction of capacity either becau se the
capacity ratio has been optimistically estimated or because of an anyhow high CAPWAP end
bearing. This problem would be aggravated if the end bearing actually relaxes. Thus,
CAPWAP results for shales or dilating soils, such as very dense fine sands or silty sands,
should never be modified analyzed with this option. Also, end bearing in cohesive materials
cannot be expected to increase significantly with large sets.
 The second extrapolation option produces conservative results and would be applicable to
sensitive cohesive (and maybe also calcareous) soils with shaft resistance reductions to a
residual value for large displacements (see Figure 4.9.1b). The user has to enter the pile
length (beginning from the pile top) over which sensitive soils exist, the percentage of
resistance loss and over what displacement distances (in quake units) it would lose this
resistance component. For piles in highly cohesive soils, it would be reasonable to apply a
30% shaft resistance reduction over a displacement distance of may be 5 quakes (n = 5). This
option is conservative and will indicate less resistance than calculated by CAPWAP.
iii. Measured Set Multiplier may be a convenient way to show a less (more) pronounced “flat spot” in
the calculated load-displacement curve by reducing (increasing) the observed set.
iv. User Defined Capacity Limit will only show the calculated load set curve to a capacity which is less
than the CAPWAP calculated capacity. It will not display an unloading curve.

(b) Earlier CAPWAP versions imposed the Dyn(amic) Toe Displacement, Elasto-Plastic Toe Resistance as
the standard for calculating the load-displacement curve. This option is still available although, since it
sometimes causes undesirable negative values or repetitive cycles in the calculated load -displacement

CAPWAP Background Report Version 2014 39


curve, it has been demoted from being the standard. It is felt that the standard monotonically increasing
toe displacement as described above yields more satisfying results. Also experience has shown that a
more sophisticated calculation of the unloading curve, which would involve the unloading quakes is, in
general, not more accurate because (a) of the uncertainty of the CS and CT values, (b) the analysis
imposes a bottom displacement boundary condition rather than top unloading (c) of the uncertain effect of
residual stresses and finally (d) even static load tests often do not produce a very reliable unloading
curve.
(c) Uplift Monotonic Changing Toe Displacement is the last applied load-displacement option. It simulates
an uplift test by applying an upward pile motion. In this case the soil resistance is the reduced shaft
resistance of the CAPWAP result. It is the users’ responsibility to specify a Friction Reduction Factor
because the shaft resistance during uplift may not be equal to the shaft resistance during compressive
loading. The CAPWAP provided default reduction factor is 0.8. For long piles this may be reasonable,
for short piles a manual check on the available soil weight should be made to ascertain that the downward
directed friction on the pile can actually be provided by the cone shaped soil weight. Care should also be
taken that end bearing is not mistakenly assigned to the shaft in the CAPWAP analysis procedure. Thus,
for uplift capacity predictions, a critical review of the CAPWAP resistance distribution results is absolutely
mandatory.
(d) Factors Directly Applied to the Load-Displacement Curves
i.Added Quake is a useful for load-displacement curve modifications where experimental evidence suggests
that the static displacements are larger than the dynamic ones (this may be frequently the case in
cohesive soils).
ii.Cut-off is sometimes an important feature since dynamically tested piles are often longer than under
service loads or during static load tests. The elastic deformation (Load)(L cutoff )/(AE) of the cut off
section of piling will then be subtracted from the calculated displacement.
iii.E-Modulus Multiplier is helpful when it is known that the static modulus is lower than the dynam ic one
(e.g., for concrete).

(e) Offset Display Options are helpful when the user wants to include in the load-displacement curve where a
capacity value would fall that conforms to certain specifications. In fact this may be important when the
failure criterion intersects with the load-displacement curve at a capacity less than CAPWAP’s predicted
ultimate capacity. The default is no offset display. The Davisson and AASHTO offsets which include in
addition to a constant offset value an elastic compression term are pre-programmed. (Note that for the
elastic compression term CAPWAP uses the dynamic elastic modulus.) The user can also define a certain
diameter ratio such as D/30 or D/10.
(f) Smooth Soil Model rounds out the otherwise sharply defined transition from elastic to plastic resistance
transition (Figure 4.9.2). This option requires a smooth factor which CAPWAP applies to the calculated
quakes. While this option causes no change in the maximum capacity value that the load-displacement
curve approaches, it may increase the displacements that would be calculated by the elasto -plastic static
soil model, since it requires a larger displacement to reach yielding in hyperbola model.
(g) Allow for Inactivated in Dynamic Analysis is an option which does not conform to the general CAWPAP
philosophy which states that only activated resistance components are shown. Utilizing this option can be
non-conservative and is discouraged.
(h) Hide Toe Load-set curve will only display the pile top load-displacement curve. This would be a
recommended option when the toe resistance is rather difficult to define (not much difference in quakes
between toe and shaft and not much difference in MQ for different toe resistance percent ages).

CAPWAP Background Report Version 2014 40


Figure 4.9.1 Extrapolation option (a) strain hardening, (b) strain
softening

Figure 4.9.2 Calculated and smoothened elasto-


plastic static load-displacement curve

For RSA and MBA the complete loading/unloading cycles are calculated for all impacts analyzed and
displayed. For RSA, if this is not desirable and only the first cycle is to be shown, then before going to output,
the RSA option can be set to zero while no other parameters should be modified.

4.10 Data Adjustments


The basic CAPWAP procedure requires adjustment of the integrated acceleration such that the final pile top
velocity is zero and displacement is equal to the observed set (inverse of blow count),. In the PDA software,
after clicking on CW the following default timing and acceleration shifts are usually taken (depending on record
features).

1. The start time of pre-compression adjustment, T1, is set to 0 ms for diesel hammers (if pre-compression
has been recognized by the program) or a short time duration before impact for other hammers.

2. The end of pre-compression adjustment time, T2, is set to the time of the first major velocity peak.

3. The pre-compression acceleration adjustment A12 is left at zero or set to a small acceleration value such
as 0.05 g’s.

4. The beginning time, T3, of the A34 acceleration shift is usually set equal to T2.

5. The end time, T4, of the A34 adjustment is usually set to T3 plus 4 ms but not more than 1.8 L/c after
maximum top velocity. In that time period some small acceleration zero shift often occurs and since it is
applied to a small time period, it practically amounts to a quick velocity shift between times T3 and T4.

CAPWAP Background Report Version 2014 41


6. The two integration constants A34 (constant velocity shift after T4) and AC (constant acceleration shift
after T2) are calculated based on the condition of zero velocity and final set at the end of the record. The
absolute value of AC is usually limited to 0.1% of maximum measured acceleration or 0.55 g’s, whichever
is lower. Similarly, the absolute value of A34 is usually limited to 2% of maximum measured acceleration or
5 g’s although the user may occasionally use somewhat higher values to achieve a satisfactory integration.

Although the closed form calculation of A34 and AC is a reasonable approach, critical inspection of the
resulting displacement curve is suggested. The displacement curve should not steeply decrease or increase
over the later portion of the record. Among the reasons for an incorrect automatic adjustment ar e (a) an
incorrect required final set, (b) a short record and therefore a velocity that should not be zero at the end of the
record and a displacement that is not yet at its final value, and (c) an incorrect adjustment over the pre -
compression time. A displacement curve that is greater (lower) than set for most of the analysis time and
sharply reduces (increases) to final set at record end may require a smaller (larger) or even negative A34
value.

Figure 4.10.1 shows both an automatically displacement curve (red curve in the bottom diagram of the top
panel) and below a displacement with a user provided 0.5 g A34 adjustment. The more horizontal behavior of
the later record portion is more reasonable. Note, that the default adjusted curve suggested that the set per
blow was actually lower than 3 mm and it is always good to check on the reliability of the blow count/set
information. Fortunately, the effect on predicted total capacity of a small adjustment error is not very severe.

Figure 4.10.1 Effect of A34 adjustment

In addition to acceleration adjustments it is necessary to check on the record proportionality. For uniform piles
with little shaft resistance close to the gage location, the records should be proportional. If not, then for
concrete piles the WS value has to be checked. For steel piles the individual force and velocity records may
suggest high bending, an incorrect calibration, a faulty transducer, a loose attachment or other measurement
problem. If there is a problem with an accelerometer indicated then it should be checked whether or not the
signal of the other accelerometer by itself gives better results. High bending may suggest material non -linearity
or non-plane cross sectional deformation and it may be necessary to reduce the calibration of the higher trace.
Such considerations should involve review of the individual force and velocity curves.

CAPWAP Background Report Version 2014 42


Obviously, adjustments to the wave speed also have to be made if the wave return from the pile toe does not
match the wave-up curve behavior at time 2L/c. Even for steel piles the wave speed is not absolutely fixed and
may actually be as much as 3% higher than the generally assumed 5,120 m/s or 16,800 ft/s.

Another check should be made on a so-called velocity shift, i.e. a phase shift between force and velocity in the
beginning of the record. VT can be used to shift the velocity curve in time backward or forward in time relative
to the force curve. This shift is usually only a fraction of a time increment long and eliminates non -monotonic
negative or positive wave-up values prior to impact. Figure 4.10.2 shows the effect of an improper VT
adjustment of a half time increment in the top graph which produces a non-monotonic positive wave up
component prior to impact. In the bottom graph (actually unadjusted, i.e., VT = 0.0) the wave up curve is
perfectly smooth through the time of the first velocity peak.

Figure 4.10.2 Improper VT adjustment (top)


and correct adjustment

4.11 Automatic Analysis Options


Ideally, CAPWAP would produce the complete calculation automatically; in that case no academic discussion
about non-uniqueness would be necessary or possible, because the program would always yield the same
result for the same data. However, there are reasons why it is not advisable to perform the analysis fully
automatically. Among these reasons are:

1. Decisions on the adjustments have to be made

2. The accuracy of all input quantities has to be reviewed even during the signal matching process

3. Non-uniformities of the pile have to be properly modeled

4. Soil properties have to be reviewed during the process

5. Different solutions have to be tried

6. Limits on certain CAPWAP variables have to be reviewed

7. The data (force and velocity versus time) may not be perfect

The AC (Automatic CAPWAP) procedure basically takes the following steps:

 It is only possible to begin the AC routine after the user has done some improvement to the match.
This assures that the routine has a good starting point.

CAPWAP Background Report Version 2014 43


 First the resistance distribution is calculated with a fixed damping based on the first 2L/c time period.
 Next damping and static capacity are exchanged for an improved match.
 Then the toe resistance parameters are varied and finally the unloading parameters, including the
quakes, are adjusted.
 This procedure is repeated to determine the best possible combination of the various paramet ers.

The AF (Automatic Friction) adjustment forms an important part of the AC routine being the basis of the
resistance distribution calculation. This procedure attempts to improve the match within the first 2L/c time
period; in general, it may not help to improve the overall match since damping, quakes and unloading
parameters may have to be adjusted to conform to the new resistance distribution.

The AT function (Automatic toe parameter search) attempts to directly calculate the main toe resistance
parameters: capacity, damping, quake, gap, unloading quake. It is a direct (not trial and error) approach and
therefore is very fast but is limited in its success rate.

The RD function (static Resistance - Damping exchange) calculates for a changed static capacity
(proportionally distributed) a set of damping factors such that the early match remains unaffected – as much
as possible. This can only be successful if the changes in static resistance are a relatively small percentage
and more than 50% change of capacity is therefore not possible during one cycle . This is a highly
recommended method to change the total capacity.

The ARD function (automatic series of RD capacity changes) is recommended to perform this analysis as a
final check on the calculated capacity. The user can specify the capacity increment and the number of steps
that should be performed.

The AQRD function (automatic RD) performs the ARD function without user interrogation, and is also highly
recommended.

The AQ procedure (Automatic Quantity search) is a grid search, i.e., it checks how small increases or
decreases of multiple user selected CAPWAP parameters would change the match quality. If improvements
can be achieved then an optimum of the quantity is sought subject to certain limits. The procedure is repeated
as long as improvements occur within a range of allowable parameters.

The AQ function Performs AQ on only the cursor selected quantity.

The AQ STD is the AQ procedure applied to a set of “Standard” CAPWAP input parameters, and is also highly
recommended. The F5 function key quickly activates this command.

All of the above automatic search tools help speed up the CAPWAP work. However, they do not relieve the
analyst from systematically investigating various capacity values and their distributions to arrive at the best
and most reasonable solution.

4.12 CAPWEAP
For driven piles it is often desirable to generate a bearing graph for an analyzed situation. A bearing graph is a
relationship between bearing capacity and blow count or set per blow. The reason for the need of a bearing
graph could be, for example, the difference between the blow count at the required ult imate capacity and the
observed blow count at the calculated CAPWAP capacity.

One way to generate a bearing graph is the so-called “Refined GRLWEAP Analysis” (Rausche et al, 2009).
Given the CAPWAP calculated capacity it requires a close matching of calculated transferred energy, pile
stresses, diesel hammer stroke and blow count with the measured values. Also, because of differences
between the CAPWAP and GRLWEAP soil models (and to lesser extent pile models), an adjustment of the
CAPWAP calculated damping factors and quakes may be needed. This process is often difficult, time
consuming and may be even frustrating.

Since CAPWAP can calculate the blow count as a function of capacity, it is possible to avoid the significant
effort of a Refined GRLWEAP analysis and instead run CAPWAP for different capacities. The hammer
performance would be automatically matched (for further thoughts on this, please see below). This procedure,
called CAPWEAP, has been programmed as an option and is accessible in the Anal ysis option menu of the
CAPWAP output program. It first involves the selection of up to 10 different capacity values and second a
choice of either standard soil model or the CAPWAP soil model. The standard soil model would include CS,

CAPWAP Background Report Version 2014 44


CT, UN values equal to 1 and neither plug nor radiation damping. Using the standard soil model would change
the calculated blow count and the match. For that reason it is not recommended to work with the standard soil
model in CAPWEAP. Also, the CAPWEAP analysis is performed for all capacities with the percentage of shaft
resistance equal to the CAPWAP calculated value while Smith damping factors would remain constant for all
capacities.

In general, CAPWAP does not produce an exact match of the measured blow cou nt and the resulting bearing
graph would therefore miss the CAPWAP capacity-measured blow count point. To avoid this confusion,
CAPWEAP optionally offers to force the bearing graph through the CAPWAP/measured blow count point by
factoring all calculated blow counts with the ratio of measured to CAPWAP calculated blow count.

Please note the following differences between GRLWEAP and CAPWEAP.

 The measured wave down curve is used in CAPWEAP as a pile top boundary condition. Strictly speaking,
this wave down curve only pertains to the situation measured and analyzed. Hammers perform differently
for different capacity levels and the wave down curve is also in part a result of the reflected wave up
curve. The CAPWEAP bearing graph should therefore only be used for capacities that are not too far away
from the calculated CAPWAP capacity, for example not more than 10%, although it is difficult to give exact
limits.

 The CAPWEAP analysis, consistent with CAPWAP, shows only activated capacities as a result.
GRLWEAP, on the other hand shows the fully analyzed capacity, even if was not activated.

As an output, CAPWEAP not only provides the bearing graph but also calculated pile top force (velocity) vs.
time obtained by adding (subtracting) calculated wave-up to (from) measured wave-down.

CAPWAP Background Report Version 2014 45


5 CAPWAP UNKNOWNS AND SENSITIVITY

5.1 CAPWAP Unknowns


The deeper the pile embedment and the more segments are needed to represent the soil resistance forces
along the pile, the greater the number of unknowns. For a pile of Np pile segments and N s soil resistance
forces (generally corresponding to an embedment of 2 N s meters) the following table gives a listing of
unknown soil resistance parameters.

Quantity Symbol Number of Type of Unknown


Unknowns
Shaft ultimate resistance Ri Ns Main Unknown
Shaft quake QS 1 Main Unknown
Shaft damping, Case (Smith) JS, (SS) 1 Main Unknown
Toe ultimate resistance Rtoe 1 Main Unknown
Toe quake QT 1 Main Unknown
Toe damping, Case (Smith) JT (ST) 1 Main Unknown
Toe, Shaft Damping Options OP, SO 2 Trimming
Shaft, toe quake multiplier CS, CT 2 Trimming
Unloading level UN 1 Trimming
Toe gap TG 1 Trimming
Shaft plug PS Np Alternate Analysis
Toe plug PL 1 Trimming
Shaft, toe radiation damper SK, BT 2 Alternate Analysis
Shaft, toe soil support masses MS, MT 2 Normally Fixed
Shaft, toe reloading LS, LT 2 Normally Fixed
Residual Stress RSA 1 Alternate Analysis

Considering usually only one shaft quake and usually only one shaft damping factor (although both can be set
individually for each of the N s elements), there are N s + 5 main unknowns, i.e. those that are of actual physical
interest and which correspond to the wave equation unknowns. Additionally, there are 7 trimming parameters
that are needed for a reasonable match, but cannot easily be associated with any meaningful soil property.
And then there are 4 options for alternate analyses which are modifying the pile model and thus require a new
start: radiation damping at shaft and toe, plug along the shaft and residual stress analysis. Associated with
the radiation damping model are the soil support masses for shaft and toe and they are usually fixed and
depend on the circumference of the pile. The reloading levels also are usually considered fixed and of little
physical significance. Furthermore, the so-called shaft plug values, which add mass to the pile segments are
seldom needed. In most instances, therefore, the number of unknowns is N s+10.

5.2 Sensitivity Study - Example Description


In order to provide the reader with a “feel” for the effect of various soil parameters and other unknowns on th e
computed pile top variables a CAPWAP analysis was performed on data taken when a 24 inch (600 mm)
octagonal pile with 15 inch (380 mm) void and a length of 94 ft (28.7 m) was driven by a (4.5 ton ram weight)
diesel hammer to 120 bl/ft (400 bl/m). The best match results are presented in the following figures in both US
and SI units. The Smith toe damping option was set to 2 (Smith damping to full activation of toe bearing then
Smith-viscous behavior afterward). For the best match, neither radiation damping nor residual stress analyses
were utilized. The final soil model parameters are as follows:

CAPWAP Background Report Version 2014 46


Shaft Damping SS s/m 0.73 s/ft 0.22
Toe Damping ST s/m 0.29 s/ft 0.09
Shaft Quake QS mm 2.5 inch 0.10
Toe Quake QT mm 10.2 inch 0.40
Toe Gap TG mm 0.18 inch 0.01
Toe Plug PL kN 0.19 kips 0.04

Unloading Level UN 0.29


Shaft quake multiplier CS 1.00
Toe quake multiplier CT 0.69

The best match quality number was MQ = 2.02 and the calculated blow count was 91 bl/ft (300 bl/m).

In the following demonstrations of the effect of variable changes on the signal match, both an increase and a
decrease of the quantity is made wherever reasonable. The resulting signal matches may then be compared
with the match on the previous page. A summary table shows the effect of the variable variation on match
quality and on blow count.

Figure 5.2.1 Final wave up match - US units

5.3 Resistance Distribution Variation


Changes in the resistance distribution (shaft and toe) have a great effect on the computed wave. In this
example 80 kips (356 kN) were taken from the toe resistance and added to element 5 as friction which was
16.9 kips before the change and became now 96.9 kips. A higher shaft resistance increases the calculated
wave-up force in the loading section of the record (time period between impact and 2L/c later), but at the same
time, the lower toe resistance reduces the calculated curve at time 2L/c. Unloading changes in the computed
wave are also evident with resistance distribution variations. The match quality number increases to 4.28 and
the calculated blow count drops to 74 bl/ft (243 bl/m). Next, the shaft resistance in element 5 which is 16.9
kips (75.2 kN) was reduced to zero while the toe resistance was increased to by 16.9 kips. The beginning of
the curve decreases due to the lower shaft resistance, but the computed wave at 2L/c shows an increase due
to the higher capacity at the toe.

RS-5 kips kN MQ Bl/ft Bl/m


Best 16.9 75.2 2.02 91 300
Low 0 0 2.12 95 311
High +80 356 4.28 74 243

CAPWAP Background Report Version 2014 47


Figure 5.3.1 80 kips (356 kN) taken from the toe and added to element 5
as friction

Figure 5.3.2 16.9 kips (75.2 kN) taken from element 5 and added to the toe
resistance

5.4 Shaft Damping Variation


The Smith Shaft (SS) damping factor for the best match is 0.223 s/ft (0.731 s/m). With a resistance of 680
kips (3025 kN) the Case Shaft (JS) damping factor is 0.429. A variation in either Case or Smith damping
factors will proportionally change the other one as long as no change in resistance is made. If either damping
factor is increased the “dynamic” resistance of the soil will increase, thus the calculated wave up in the first
2L/c will also increase. In the same way, this increase will dampen the signal at a fast er rate, thus the “valley”
at 2L/c will be filled in and the unloading part of the wave will decrease with th is change. On the other hand a
decrease in the Smith damping factor will decrease the wave up signal in the 2L/c time range, but the
unloading part of the signal will increase..

Shaft Damping
Variation MQ s/ft s/m Bl/ft Bl/m
Best 2.02 0.223 0.73 91 300
Low 5.9 0.10 0.31 71 233
High 5.7 0.40 1.3 120 393

CAPWAP Background Report Version 2014 48


Figure 5.4.2 Smith damping value increase

Figure 5.4.1 Smith damping value decrease

5.5 Toe Damping Variation


Variations of the toe damping factor will only take effect when the wave arrives to the bottom of the pile and
will have an effect on the wave-up match only L/c later. Thus the initial part of the match, prior to the “valley” is
not affected. Like the shaft damping, Smith and Case Toe (ST and JT) damping factors are directly related if
no change to the static resistance is made. An increase to the Smith Toe damping from the best match value
of 0.088 s/ft (0.289 s/m) to 0.3 s/ft (0.984 s/m) will increase the computed wave up at time 2L/c but the
unloading part of the signal will decrease. A decrease in toe damping will cause a reduc tion of the wave up
signal at 2L/c and a slight increase of the unloading part of the curve.

Toe Damping
Variation MQ S/ft S/m Bl/ft Bl/m
Best 2.02 0.088 0.29 91 300
Low 7.6 0.01 3.0 84 274
High 5.9 0.30 1.0 150 493

Figure 5.5.1 Smith toe damping value increase

CAPWAP Background Report Version 2014 49


Figure 5.5.2 Smith toe damping value decrease

5.6 Shaft Quake Variation


The Skin Quake (QS) for the best match CAPWAP analysis has a value of 0.1 inch (2.54 mm) as is commonly
experienced. For a fixed capacity, lower quakes will quickly activate the resistance, thus the force in the wave -
up in the first 2 L/C time period after impact will increase while a larger quake will cause it to decrease. The
opposite effect would occur in the unloading part of the curve because resistance is more quickly lost for the
lower quake. The unloading effect can be offset by a change in CS value which would then modify the
unloading quake. However, there is a limitation on the unloading quake: in general, it cannot be greater than
the loading quake with the exception of situations that involve the radiation damping analysis.

Skin Quake
Variation MQ inch mm Bl/ft Bl/m
Best 2.02 0.10 2.54 91 300
Low 9.3 0.01 0.25 125 404
High 4.2 0.40 10.2 179 583

Figure 5.6.1 Skin quake increase

Figure 5.6.2 Skin quake decrease

CAPWAP Background Report Version 2014 50


5.7 Toe Quake Variation
Toe Quake (QT) effects take place as the compressive wave reaches the bottom of the pile. Similar to the
shaft quake effect, a toe quake increase will decrease the wave-up force (or delay the effect of the end
bearing) at time 2L/C followed by an increase in the unloading part of the wave -up. In the present example the
decrease of wave-up is particularly obvious because both static resistance and damping are activated later
(OP is 2 and therefore damping is linked to static resistance) by the larger toe quake. A decrease of the toe
quake will have the opposite effect. A value of 0.4 inches (10.16 mm) was estimated for this example’s best
match.

Toe Quake
Variation MQ inch mm Bl/ft Bl/m
Best 2.02 0.4 10.2 91 300
Low 8.7 0.1 2.5 72 237
High 8 0.8 20.3 63 205

Figure 5.7.1 Toe quake increase

Figure 5.7.2 Toe quake decrease

5.8 Unloading Limit Variation


Variation of the Unloading (UN) limit variable will only affect the match when the pile velocities become
negative, in this case, after the 2 L/C wave travel time. UN limits how much negative (downward directed)
shaft resistance can act along the pile during pile rebound. The variation of this parameter therefore has an
inverse effect on the late portion of the wave-up force. In other words, the late calculated wave-up will
increase if the parameter is decreased, or will decrease if the value is increased. The current best match
number for UN is 0.29 meaning that the shaft resistance is limited to -29% of the positive ultimate resistance
value during rebound. This variable also influences the calculated blow count, since the end part of the record
determines the permanent set of the pile.

CAPWAP Background Report Version 2014 51


Unloading Limit
Variation MQ Bl/ft Bl/m
Best 2.02 0.289 91 300
Low 2.4 0.1 136 446
High 7.0 1.0 47 155

Figure 5.8.2 Unloading limit decrease

Figure 5.8.1 Unloading limit increase

5.9 Toe Gap Variation


When driving a pile, bouncing can generate a small gap between the bottom of the pile and the underlying soil
or rock. Under the next blow, the pile toe will move a distance through the gap without resistance until the gap
is closed and the resistance is engaged. Sometimes strain hardening occurs, i.e. the end bearing is at first soft
and then starts to gain stiffness, for example due to a negative pore water pressure build up. These situations
can be modeled with a Toe Gap (TG) which will affect the computed wave-up curve only at the 2 L/C time, i.e.
when the wave returns to the pile top after its first toe reflection. In the present case TG was practically zero,
0.01 inches (0.02 mm) for a best match and could not be much further reduced. A small and a large increase
of TG was therefore chosen for this study. Note that the gap effect would be further amplified if the toe
damping option OP were set to either 1 or 2 such that damping would be zero as long as the toe resistance is
zero.

Toe Gap Blow Count


Variation MQ inch mm Bl/ft Bl/m
Best 2.02 0.01 0.18 91 300
Small 4.6 0.10 2.5 77 252
Large 13.4 0.30 7.6 39 129

CAPWAP Background Report Version 2014 52


Figure 5.9.1 Small toe gap increase

Figure 5.9.2 Large toe gap increase

5.10 Unloading Shaft Quake Variation


CS (the ratio of the unloading skin quake to the loading shaft quake) does not affect the computed signal in
the beginning of the record, but rather only after the pile velocity is negative, or rebounding. In the example,
the best match value of CS was 1.0 which means loading and unloading quakes are equal. An increase
beyond 1.0 is only reasonable in a radiation damping analysis. A decrease of this parameter will cause static
resistance to decrease quicker during unloading (rebound) and therefore lower the l ater unloading section of
the computed wave. In general, skin unloading variations are expected prior to the UN effects. In addition, the
timing of the CS effect is dependent on the energy that has been transferred to the pile and the soil resistance
at both shaft and toe.

CS Blow Count
Variation MQ Bl/ft Bl/m
Best 2.02 1.0 91 300
Low 3.40 0.2 100 329

Figure 5.10.1 CS decrease (in general CS should not be lower


than 0.3)

5.11 Unloading Toe Quake Variation


CT is the ratio of unloading toe quake to loading toe quake. As discussed for CS, CT affects the unloading
portion of the match, i.e. after the time of maximum toe displacement. This is indicated on the screen by the

CAPWAP Background Report Version 2014 53


last upward tick mark (dashed) below the time axis. Normally CT’s range is between 0.3 and 1, however for
large loading quakes, like the 0.4 inches (10 mm) in the present case, lower values of CT are also reasonable.
Higher values than 1.0 are possible when there is a Toe Gap greater than 0 or when radiation damping is
employed. A low CT would unload the toe resistance quickly and therefore lower the later part of the
calculated wave-up curve. In the following example, the low value would make the effective unloading quake
0.4 x 0.1 = 0.04 inches (1 mm).

CT Blow Count
Variation MQ Bl/ft Bl/m
Best 2.02 0.69 91 300
Low 7.7 0.10 51 168
High 3.0 1.0 157 515

Figure 5.11.1 CT decrease

Figure 5.11.2 CT increase

5.12 Toe Plug Variation


A soil plug mass at or underneath the pile toe will have an inertia effect. Its mass can be specified through its
weight input, PL; it will affect the calculated wave around the 2L/c valley. A plug mass increase will produce a
positive effect (positive acceleration produces a compressive wave) just before 2L/c and opposite negative
effect immediately following 2L/c (negative acceleration produces a tension wave). Thus, a phase shift of the
calculated wave-up curve from left to right is expected when plug mass is increased. In this case study, the
plug for the best match has a value of 0.04 kips (0.19 kN). The resulting match appearance is therefore one of
a somewhat longer pile. It should be noted that an excessively high plug weight would cause an effect similar
to high toe damping. In the variation example, since the best match plug weight was very small, reducing it did
not produce a clear effect.

Plug Blow Count


Variation MQ kips kN Bl/ft Bl/m
Best 2.02 0.04 0.19 91 300
Low 2.07 0.00 0.00 90 297
High 6.11 2.00 8.89 84 2.75

CAPWAP Background Report Version 2014 54


Figure 5.12.2 Plug weight increase

Figure 5.12.1 Plug weight decrease

5.13 Shaft Plug Variation


A soil mass can also be specified for the shaft segments. This may be helpful for plugged pipes or H -piles or
for drilled piles with a rough surface which causes a shear failure some distance away from the pile concrete
and therefore causes soil to move with the pile shaft. The first segment cannot be varied. The PS input is in
force units (kips, kN) per segment. In the current example, the individual segment weights were roughl y 2 kips
(9 kN). If the 15 inch (380 mm) diameter void of this concrete pile were fully filled with soil then the added
weight at each pile segment would be roughly 0.3 kips (1.4 kN). It is recommended, however, to input values
that are less than a segment weight. Greater PS values could cause numerical instability. PS will affect the
match where the segment weights increase and will also produce the effect of a reduced wave speed. In the
example below, a plug mass of 0.2 kips (0.9 kN) was added (a) to all segments and (b) only to the lower half of
the pile. Note that the top segment properties cannot be modified by plug mass since the plug mass causes an
impedance increase.

CAPWAP Background Report Version 2014 55


500.0 kips
W up Msd
W up Cpt

250.0

14 74 ms
0.0
7 L/c

-250.0

Figure 5.13.1 PS = 0.2 kips (0.9kN) - segment 2 to toe

500.0 kips
W up Msd
W up Cpt

250.0

14 74 ms
0.0
7 L/c

-250.0

Figure 5.13.2 PS = 0.2 kips (0.9 kN) - lower half of pile

5.14 Radiation Damping Variation


SK and BT are both radiation damping parameters. Zero values are interpreted as an infinitely stiff soil
surrounding the pile (which is equivalent to the standard Smith soil model). High numbers (e.g. 100) are also
equivalent to a very rigid shaft and toe soil support and would therefore show little effect on the calculated
quantities. Specifying a SK or BT value greater than zero will automatically assign a reasonable, pile size
(perimeter and pile end bearing area) dependent MS or MT soil support mass between the shaft or toe
resistance and soil support dashpot. The use of recommended values is strongly encouraged if radiation
damping is utilized. Small SK or BT numbers should be used with extreme caution. Small values under this
recommended range can lead to great variations in the wave-up force, and an unconservative
overprediction of capacity and thus should be avoided. SK could affect the computed wave-up from the
beginning of the record, and BT only affects the record after the wave returns from the toe. While some limited
experience exists with SK (Likins et al., 1996), the bottom damper, BT, recommendations are based on
intuition rather than correlation and should therefore be used with great caution.

The best match of this example had an infinite shaft radiation damper (SK = 0). The example shows that
positive SK values, and therefore a softer soil response causes a quick unloading of the shaft resistance and
therefore a reduced wave-up already immediately before the “valley”, because unloading begins al ready while
the pile is still moving downward. Reviewing the background information and the paper by Likins et al., (1996)
and besides performing a thorough sensitivity study on the effect of the radiation damping model is strongly
recommended before using this analysis method.

The bottom radiation damping causes a similar early unloading effect as SK, however it begins later.

CAPWAP Background Report Version 2014 56


Figure 5.14.1 SK = 0.35

Figure 5.14.2 SK = 0.70

Figure 5.14.3 BT = 1.0

CAPWAP Background Report Version 2014 57


Figure 5.14.4 BT = 5.0

5.15 Toe Damping Option Variation


Three options are available under the Toe Damping Option OP:
0-Viscous,
1-Smith,
2-Smith to full static resistance activation, then viscous.
The present best match was obtained with the use of the Smith-viscous damping approach for the toe (OP =
2). OP = 1 and 2 will produce a similar match around the 2L/c valley where their damping effect will increase
slower if the toe quake is large. OP 0 will produce higher damping effects at time 2L/c because the damping
factor is unaffected by the static resistance activation. The unloading section of the computed curve will be
differently affected by all three options. Increasing OP to 1 from 0 will use the traditional Smith damping
approach which provides very little negative damping effect when static resistance components are small and
velocities negative.

500.0 kips
W up Msd
W up Cpt

250.0

14 74 ms
0.0

-250.0

Figure 5.15.1 OP = 0
500.0 kips
W up Msd
W up Cpt

250.0

14 74 ms
0.0

-250.0

Figure 5.15.2 OP = 1

CAPWAP Background Report Version 2014 58


5.16 Shaft Damping Option Variation
Three options are available under the Shaft Damping Option SO and are equivalent to OP.
0-Viscous,
1-Smith,
2-Smith to full static resistance activation, then viscous.
Different from OP, however, the effect of these options will be apparent already in the first 2L/c time period of
the match. The best match had been obtained with the SO=0 option (standard). The effect of the SO=1 and 2
options is shown below. Note the much stronger damping effect (similar to SO=0) of SO=2 compared to SO =1
over the later part of the match.

500.0 kips
W up Msd
W up Cpt

250.0

14 74 ms
0.0

-250.0

Figure 5.16.1 SO = 1

500.0 kips
W up Msd
W up Cpt

250.0

14 74 ms
0.0

-250.0

Figure 5.16.2 SO = 2

CAPWAP Background Report Version 2014 59


5.17 Help
The response of the Wave-Up match to some of the CAPWAP variables is graphically depicted in the HC help. This
help shows the current match and divides it into certain time periods which are outlined by arrow lines. Above these
lines are shown in red the names of those parameters that when increased would cause an increase in the calculated
wave up. Underneath the line, in blue, the name of the parameter is shown that, upon increasing it would cause a
decreased wave up. For example, in the first 2L/c portion of the match, increasing RS (static shaft resistance) or JS
(shaft damping factor) would cause an increased wave up. On the other hand, if the QS (shaft quake) value were
increased then the wave up would be lower in that segment of the match.

Figure 5.17.1 The HC Help Screen

Figure 5.17.2 The HR Help Screen

The CAPWAP Resistance Screen, HR, (Figure 5.17.2) may also be helpful for an understanding of the resistance
development at one user selected segment. The screen shows the relative magnitude of static and dynamic
resistance forces as a function of segment displacement, on the left, and the relative magnitude of loading and
unloading quakes together with the reloading level on the right.

CAPWAP Background Report Version 2014 60


6 CAPWAP PROCEDURE

6.1 Introduction
The CAPWAP procedure should not be a blind minimization of the differences between measured and computed
signals. Instead for every step taken in the matching process the analyst should be cognizant of the implications of
the parameter choices made. The lowest match quality number is only then associated with an acceptable solution if
the selected model parameters establish a physically meaningful combination. For example, very high damping
factors may be reasonable when pile velocities are unusually low (low impact velocity, relatively high resistance) or
the soil is a highly cohesive clay or silty clay. On the other hand a high damping factor is hard to explain when a pile
is driven to rock.

The automatic routine, AC, in general follows the recommended signal matching procedure, however, it should only
be employed once a reasonably good solution has been achieved. It then can provide an independent check on the
results achieved. Finally, since the CAPWAP analysis involves non-linear soil behavior and a relatively large number
of unknowns, it is important to try different sets of parameters (e.g. relative amount of end bearing versus shaft
resistance or damping factors or quakes) as starting points before accepting a final solution. In addition, before the
final solution is accepted, both higher and lower capacity values (use RD procedure, or AQRD) should be tried to
obtain the most realistic and best match (lowest MQ) solution.

6.2 Unknowns
The normal matching procedure involves the following unknowns (Ns = number of shaft soil resistance segments):

Ns ultimate shaft resistance values R(1) ... R(Ns),


1 ultimate toe resistance value, R(Ns+1),
1 shaft loading quake (QS) plus 1 shaft unloading quake multiplier (CS),
1 toe loading quake (QT) plus 1 toe unloading quake multiplier (CT),
1 shaft damping factor (SS),
1 toe damping factor (ST),
1 unloading level (UN),
1 toe gap (TG),
1 toe damping option (OP),
1 shaft damping option (SO)
1 plug mass at the pile toe (PL).

Thus there are Ns+12 unknowns in the normal case. Of course, choosing individual variable quakes, individual added
damping or multipliers, one or two added toe resistance values, different reloading levels, radiation damping, shaft
plug, shaft damping options and the RSA option can greatly increase the number of unknowns. Because of the
increased associated uncertainty, these options should only be chosen once all other possible solutions have been
investigated.

6.3 Analysis Steps


The following steps have to be taken when performing an analysis

• Record selection
• Data check and adjustment
• Pile modeling
• Signal matching and best match generation
• Output generation
• Result interpretation

Each step requires careful review and analysis. Because of the great variety of hammers and their performance
characteristics, pile sizes and materials, pile lengths, pile installation procedures, soil types, time and type of testing
only a general frame work for a successful analysis can be outlined. Indeed, every case analyzed may provide the
analyst with a new learning experience. The following gives a general review of the necessary analysis steps; more
detailed information is included in the HELP section of the program.

6.4 Record Selection


Arguably, record selection is the most important part of the CAPWAP analysis: if the record is from a blow with
questionable data quality (excessive bending, loose attachment bolts, incorrect calibrations, intermittent wire failure,
etc.), low energy in hard driving, high energy in easy driving, late during restrike or too early prior to the end of

CAPWAP Background Report Version 2014 61


driving, CAPWAP results may not yield the intended result. The following are simple guide lines for record selection,
but the analyst is also encouraged to use best judgment since no set of guide lines can be complete or perfect:

• If the average set per blow is less than 3 mm (blow count more than 100 blows/ft), select a blow with high
energy and force level to avoid under-prediction.
• If the average set per blow is greater than 12 mm (blow count less than 25 blows/ft), select a low energy blow
to avoid over-prediction. While not impossible to analyze such a blow, the difficulty in matching is certainly
increased.
• For sets between 3 and 10 mm/blow (30 to 100 blows/ft), select a record with average energy for an average
capacity result.
• Select a record with minimal bending effects (check individual strain records), no spikes or other electronic
noise, integrating out to a zero velocity and a final displacement equal to the observed permanent set, with
zero or slightly negative final force and good proportionality (if the pile is uniform and the distance between
sensors and shaft resistance effects substantial). If the force at the end of the record is positive (and the
velocity has clearly stabilized to zero) there may be plastic yielding of the pile material and this data must be
rejected since the pile is no longer linear elastic.
• In a restrike situation an early record should be chosen; if the early records are of low energy and the set per
blow is very low, then it is important to find a good compromise between when the record occurs and its
energy. The “first high-energy blow” is then preferred. If the record was taken late in the restrike and the soil is
easily losing resistance then some of the setup capacity may have been lost prior to the record analyzed (see
Section 7.1 for description of the superposition procedure). It is also possible to analyze more than one blow
and superimpose results (if and only if the set is at refusal).
• For end of driving, find a late record.

In any event, the selection of the record with the best data quality is very important, selecting the wrong record could
add hours of extra effort and then you still may not be able to achieve a good match. Even if you did get a good
match, the answer may be unreliable. The reasons for bad data quality may include:

• Electronic noise, e.g. due to wet or damaged cables or neighboring powerful transmitters or power lines
• Distortion of the force record due to extreme pile top bending or damage.
• Distortion of the force record due to local bending or shear effects. Stay away from welds in steel piles or
from joints of cracks in concrete piles. Try to also keep away from cross section area changes.
• Distortion of the force or acceleration signals due to slipping transducers.
• Accelerometer overloads (e.g. for uncushioned impacts or those with high lateral acceleration components;
overloads may cause a sharp decline of the velocity after the first peak.
• Wrong calibration input of strain transducers or accelerometers.

The record selection therefore should be based on the following assessments of data quality:

• Good force-velocity proportionality, unless pile non-uniformity or friction acting near the sensor location dictate
otherwise
• Signals without spikes or excessive high frequency noise
• A velocity record that exhibits at the end a stable zero amplitude or an oscillation around the zero line
• A final pile displacement that corresponds to the observed blow count (if it is accurately known and the record
is long enough). If velocity is not at zero at the end of the record a reasonable adjustment should be made or
better yet, if the signal looks reasonable and the accelerometer is functioning properly, no adjustment should
be made. The adjustment may have to be revised during the matching process. (Note: if the final
displacement is not equal to the observed set or if the observed set is at best estimated, blow count matching
should not be attempted. The observed set for the analyzed blow may also be adjusted, particularly in
restrike, if there is evidence of a change in hammer performance for different blows or if the soil is sensitive
and loses resistance blow by blow; judgment then may guide what the appropriate set might be).
• Strain records that return to the zero line (or slightly negative) at the end of the record (no pile material
yielding or strain record distortion due to concrete cracking under the transducer or no transducer slippage).

Sometimes several records have to be analyzed, particularly in the earlier mentioned case of restrikes where
simultaneous energy increases and shaft resistance decreases may require application of the method of
superposition (see Section 7.1).

Averaging of several records is possible when record selection is done, however, it is only recommended for
situations where random noise seems to make for a rough appearance of the signals and where consecutive hammer
blows had similar energies and soil resistance was practically constant. To accomplish analyzing the average of more
than one blow:

CAPWAP Background Report Version 2014 62


 Locate first blow in the PDA program and without closing PDA, click back to CAPWAP and select ADD
RECORD
 Click back to the PDA program, find the next blow of interest and without closing PDA, click back to CAPWAP
and select ADD RECORD
 Continue adding all desired blows in a similar manner, and when the list of blows is complete, click OK in the
“Get data from PDA” dialog box.

6.5 Data Adjustment


Because of the imperfect nature of all measurements and the need of accurate data in CAPWAP, it is necessary to
check, and, if possible and necessary, adjust the measured force and/or velocity data. Other data adjustment
parameters are calibration factors, smoothing or filtering, time shifts and amplitude shifts. Three different adjustments
have to be considered:

• Calibration adjustments of a few percentage points are not unreasonable, considering that the accelerometer
and strain transducer calibrations are generally only accurate within 2 percent and that material properties of
concrete and even steel also tend to introduce a calibration error in the force calibration.
• Integration of acceleration over relatively long times can accumulate so-called acceleration zero shift errors.
Thus slight shifts of the acceleration zero line are necessary (see Section 4.10). The acceleration shifts are
probably the most complex adjustments, may have some consequences on the matching process and thus
the capacity calculation and are relatively frequently applied. However, they are important, particularly when
blow count is to be matched. The absolute magnitude of these adjustments is very small compared to the
maximum acceleration levels that are typically measured near the pile top.
• A less important, more cosmetic adjustment, is a time or phase shift of velocity relative to force (VT in PDA
program). Such an adjustment is normally only a fraction of a time increment. It should be chosen such that
the wave-up curve is smooth at the time of first velocity impact peak.

It is sometimes reasonable to begin the signal matching analysis without prior data adjustment and only once it has
become clear that a good match cannot be achieved without data adjustment, should the necessary adjustments be
made. The Default adjustment in the PDA program normally yields very reasonable results.

6.6 Pile Model


For uniform piles the program automatically generates a model that assumes constant pile properties from top to
bottom. If the pile is non-uniform the user has to modify this assumed uniform model in PM prior to signal matching
work. The known non-uniform properties usually only involve the cross sectional area. Sometimes also the pile
material also changes and then elastic modulus and specific weight might have to be changed. These are the
standard pile profile parameters. Also the perimeter value may have to be changed for a correct calculation of the unit
resistance values in the output. The end bearing (or toe) area, responsible for a correct unit toe resistance) also may
need a correction. Non-standard model adjustments have to be done in an iterative manner during the matching
process (see Section 6.9) they may involve:

• an overall wave speed adjustment (assuming that the pile is of concrete and that its properties at the top differ
from those at other pile locations.
• an impedance variation, e.g., when a drilled shaft is analyzed or an open ended pipe is plugged which adds
mass and stiffness.
• a soil plug along the shaft, for example to add the weight of soil in an open ended pipe or between the flanges
of an H-pile. (usually for only a few pile elements and rarely if ever for the entire pile length)
• tension slacks for situations where cracks have formed and the return wave has been lengthened by a tension
wave opening these cracks.
• compression slacks when there is an obvious early tensile reflection (prior to the return of the wave reflected
from the pile toe) which cannot be matched with an impedance reduction.
• very rarely a pile damping variation since its default is reasonable and generally accepted as a standard
parameter.
• also seldom required is a time increment change (see 6.9.5)

Always, try first matching without these pile model features and only if a good match seems impossible, try these
adjustments. For non-uniform piles it is important to enter the so-called pile profile, i.e. the variation of cross sectional
area, elastic modulus, specific weight and perimeter before attempting any other pile model adjustments.

CAPWAP Background Report Version 2014 63


6.7 Signal Matching
After data input, record adjustment and pile model setup has been completed, the actual signal matching process
begins. The process must proceed from “left to right”, i.e. beginning at the time of impact through the end of time in
the record. For the very first match, CAPWAP assigns a capacity value based on the Case Method and a resistance
distribution based on the appearance of the wave-up record or, for short piles a triangular distribution. Also, CAPWAP
assigns standard wave equation recommendations to damping and quake parameters. The user then should
complete the following tasks:

I. For non-uniform piles, enter in the pile model (PM) the non-uniform pile profile, if it is known (for drilled shafts,
the area versus length may be variable and unknown and requires judgment by the experienced analyst to
consider the concrete installation logs as well as the soil profile to find a reasonable solution.)
II. Improve the match over the first 2L/c time period by varying individual resistance values and total capacity. As
a help in the resistance input, the Delta value indicates the difference between computed and measured pile
top quantity at a distance 2xRi/c where the resistance i has its effect on the match.
III. Improve the match at and immediately after the 2L/c reflection by modifying the end bearing and the toe
quake (and/or toe gap) with simultaneous total capacity change and variation of toe quake and toe damping.
IV. Improve the match after 2L/c by modifying total capacity with simultaneous adjustment of shaft and toe
damping and then repeat the process by going back to step b.
V. Improve the match at 2L/c by trying different toe damping options using OP. (Usually the OP is either 0 or 2).
Go to step c.
VI. Improve the late record match by modifying unloading quakes and unloading parameters. Since this step may
also affect the early record portion, restart at b.
VII. Vary total capacity, restart at b.
VIII. Review whether or not pile model changes are necessary and, if so, restart at a.

In the following examples, computed and measured force curves are compared and the necessary adjustments to the
individual Ri values are discussed. Conclusions would be similar for wave-up comparisons except that differences
between the measured and computed wave-up curves would only be ½ of the differences between computed and
measured quantities. For velocity matching, the sign of the differences would be reversed and since velocity is
displayed after multiplication by Z, the differences would be shown in force units.

Example 1. Lack of Resistance


If the measured and computed force (wave-up) curves are parallel as shown in Figure 6.7.1 between elements 3 and
Ns, then the total resistance for those elements was correctly determined. However, the Delta value “d” corresponds
to the resistance value that must be added over the first three elements where the measured and computed force
curves diverge. Sufficient accuracy is probably obtained by increasing the static resistance values at segments 1, 2

Figure 6.7.1 Force match showing lack of static resistance at top 3


segments

and 3 by roughly d/3 (2d/3 if these are wave-up curves).

Example 2. Excess Resistance


In the example of Figure 6.7.2, too much resistance acts on elements 6 and 7. Reduce both R6 and R7 by d/2 (d in
the case of wave-up matching).

CAPWAP Background Report Version 2014 64


Figure 6.7.2 Excessive static resistance

Example 3. Localized Lack of Resistance


In the case of Figure 6.7.3, there is not enough resistance between elements 4 and 6. Thus, d 1/3 (2d1/3 for wave
matching) needs to be added to RS(4), RS(5) and RS(6). Adding resistance to these segments will, of course, raise
the computed curve in this region, but it will also raise the curve later. There is already too much resistance at
elements 7 through 11 and (d1 + d2)/5 must be subtracted from soil segments 7 through 11 [2(d 1 + d2)/5 for wave-up
matching).

Figure 6.7.3 Force match showing low resistance at segments 4 to 6 and


high resistance below that

Example 4. Low Static Capacity


In Figure 6.7.4, the computed force is low after 2L/c. Capacity of an amount approximately equal to ½ the Delta
value (equal to the wave-up difference) must be added and a corresponding amount of damping should be taken
from the shaft. Alternatively, the end bearing could be increased. Merely adding static capacity to all soil segments
would destroy the match in the early portion of the record.

Figure 6.7.4 Force match showing good shaft resistance match, but low
total static resistance

Example 5. High Static Capacity


In Figure 6.7.5 the computed force is high, therefore, the static capacity must be reduced. Since the match begins to
diverge at L/c, either reductions of the resistance values in the lower half of the pile are advised or all Ri values
should be proportionally reduced and an equivalent amount of soil damping added.

Since soil damping is calculated as the product of segment velocity times damping factor it is highest at a certain
segment, when the peak of the impact wave passes that segment. Damping resistance, therefore, makes for a high
resistance effect only for a relatively short time compared to static resistance. Damping forces also effectively smooth
high frequency mismatches and oscillations in the computed curve. The best indication of the need for damping is

CAPWAP Background Report Version 2014 65


found in the portion of the record immediately following the time 2L/c after impact: if too much damping and not
enough static resistance has been used to match the first 2L/c period then the total capacity will be low which makes
the computed force (or wave-up) low during the time after 2L/c.

Figure 6.7.5 Force match showing excessive static resistance

Figure 6.7.6 Force match showing evidence of insufficient damping

Example 6. Insufficient Damping


In Figure 6.7.6 the computed force curve has large oscillations after 2L/c. Add soil damping to smooth the computed
force curve (as a shock absorber would smooth out the ride in a car). Since the match is actually satisfactory prior to
2L/c, damping should be added to the toe. If that is not possible (toe damping is getting too high) then damping
should be added to the shaft which requires reducing the shaft resistance values, R i, and which in turn will require an
increase in Rtoe. The RD and AF routines are very helpful in that process.

Example 7. Non-proportional Damping


Figure 6.7.7 shows a local mismatch at segments 3 and 4. Let us assume that this match did not involve much static
resistance below No. 6 that could be shifted upwards for a better match. Then the preferred and probably only
successful solution to improve the match is to use extra damping at segments 3 and 4 (perhaps this corresponds to a
clay layer). Static resistance at 3 and/or 4 would produce a high computed curve at and below 5 and would not be
satisfactory. Added Damping, Damping Multipliers, Impedance or Plug on Shaft could be used to improve the
match. These tools provide a localized effect on the computed quantity.

It is always important to find out why such parameters should be chosen. Impedance changes would be reasonable
where a high temporary plug adds stiffness and weight and the Plug on Shaft if the plug only adds weight. Note that
damping in the presence of zero static resistance leads to an “infinite” Smith damping factor. This added damping
could be explained by a rather viscous upper soil layer that has not much static resistance. Extra damping is also
indicated for modeling friction effects of template guides, i.e., where unusually low (or zero) static resistance values
are accompanied by significant dynamic resistance values. The damping multipliers are particularly useful when it is
known that a particular soil layer has high damping relative to other layers.

CAPWAP Background Report Version 2014 66


Figure 6.7.7 Force match indicating need for non-proportional (added)
damping, increased damping multiplier, added impedance or plug on shaft

Example 8. (Unloading) Quake Adjustment for Late Record Match


Figure 6.7.8 shows a high computed force in the late record portion. To improve the match, quakes may be
decreased. Of course, this will cause a relatively early loading of R sk which may be compensated for either by
reducing damping or by downshifting static resistance. It would be simpler to reduce the unloading quake multipliers,
CS or CT (if that is an option, i.e. if they are not already set to low values), or increase the unloading level UN (if that
is an option, i.e., if UN is not already at 1.0). Another measure would be trying a decrease in the static capacity or
changing the damping options OP or SO.

Figure 6.7.8 Force match showing potential need for (unloading) quake
decrease

Example 9. Quake Adjustment to Delay Resistance Effect


In Figure 6.7.9, the toe resistance indicator (A) shows full activation near 2L/c. If the toe resistance could be delayed
until time (B), the match near 2L/c would be improved. A larger toe quake, QT, can create this result. (In many
instances a toe gap, TG, and/or the Smith toe damping option, OP=1 or 2 may also be necessary to improve this
match type; see also Section 6.8 and Example 10).

Notes: A decrease in toe quake is necessary if the maximum toe displacement is less than the quake. The only
alternative would be a reduction of either static or dynamic resistance for greater pile displacements. The computed
pile capacity may be in error if, at the same segment, the quake is greater than the pile displacements (non-activation
of static resistance occurs).

Figure 6.7.9 Force match showing the need for toe quake increase to delay
toe resistance effect

CAPWAP Background Report Version 2014 67


6.8 Signal Matching Using Soil Model Extensions
6.8.1 Unloading Parameters CS, CT, UN
Example 8 already pointed out that the late record may be difficult to match with shaft and toe quake changes since
that would also affect the early record portion. Quakes influence the late record primarily because they affect the
speed of unloading. A faster shedding of load (unloading) is indicated by a low computed force at the end of the
record. If and when a reduction of the loading quakes is undesirable (e.g., due to need for a slow loading rate) then
unloading can be made to act faster by reducing the unloading quakes which is accomplished by reducing the
unloading quake multipliers, CS and CT.

The unloading level parameter, UN, changes the limit of negative shaft resistance values which can occur during
pile rebound. UN = 0 means no downward directed friction is possible during pile rebound while for UN = 1 the
negative friction limit equals the positive ultimate shaft resistance. A UN of 0 value would, therefore, prevent late
negative static resistance values and therefore prevent a low computed force or wave-up force in the late record
portion.

Example 10. Matching of the Unloading Phase

Figure 6.8.1 shows an example which would require a reduced UN (say 0 to 0.3 instead of 1.0). Increased unloading
quakes may have similar effects in the late record portion, however, UN acts later than CS (during pile rebound, first
resistance is unloaded from the positive limit to zero according to CS and only then the negative friction values occur
if allowed by UN.

Figure 6.8.1 Force match with low computed force due to either high UN
value or low CS or CT

6.8.2 Reloading Levels, LS, LT


If the unloading quakes are lower than the loading quakes and if the pile exhibits a vibration like down - up - down
movement then there may be a problem with resistance activation as shown in Figure 6.8.2. This problem can be
eliminated by setting the reloading level higher (e.g., from -1 to 0.5 for the shaft or from 0 to 0.5 for the toe). The
effect is demonstrated in Figure 6.8.3.

CAPWAP Background Report Version 2014 68


Figure 6.8.2 Reloading along slope given by loading quake (low
reloading level)

Figure 6.8.3 Reloading with high reloading level

The reloading levels whose default is 1.0 (reloading follows the unloading quake path which means a very quick
reloading and, therefore, activation as for the case with unloading quake = loading quake) are non-standard
parameters and only rarely need adjustment. In that case they may be adjusted to LS = -1, -0.5, 0, 0.5 and LT = 0,
0.5; further refinement seems unnecessary.

6.8.3 Toe Gap, TG


When a pile with low skin friction is driven to a stiff layer, it often rebounds so strongly that the pile bottom loses
contact with the bearing layer. Under the next blow, the pile will move a distance (a gap distance) without
appreciable end bearing until the pile bottom contacts the bearing layer. The toe gap (TG) adds to the toe quake and,
like the toe quake, helps where the computed force curve is high at the 2L/c time (see Example 9). Of course, a toe
gap is only effective in cases with significant toe resistance, and is also more likely in cases of a large toe quake.
Note that for the final simulated static load test analysis the toe gap is combined into the toe quake. The assumption
is that the gap effect occurs only during pile driving but not in a static condition.

CAPWAP Background Report Version 2014 69


6.8.4 Damping Options OP, SOSuppose the pile toe is not in contact with hard soil and a toe gap is needed to
match a low force valley at 2L/c. The soil has the potential for large damping once pile toe and soil come into
contact. In this case the usual viscous toe damping model is inadequate because high damping forces would result
before the toe actually contacts the soil. In this case, a better approach is to follow Smith's recommendation of
multiplying the product of the viscous damping factor, Jt, and pile toe velocity, vt, with the ratio of activated to ultimate
resistance. Thus, as long as the static resistance is low, the damping resistance will also be low. Figure 6.7.9
demonstrates a case where a large quake, a gap and/or the Smith toe damping option would be helpful. In most
cases, however, it is best to switch back from traditional Smith (OP=1) to the viscous approach once the initial loading
phase has produced full resistance activation. This is done by choosing (OP=2). The reason is that in the late portion
of the record more damping is usually needed than provided by the pure Smith approach. For the shaft the equivalent
damping options SO=1 and SO=2 are much less frequently needed or useful possibly because the shaft resistance is
not allowing for a separation like the end bearing does.

6.8.5 Radiation Damping, SK, BT


When a pile is driven to rock, a wave is induced in the rock which makes the resistance vs pile bottom displacement
behavior very different from the elasto-plastic Smith approach. One characteristic of such a resistance-displacement
curve is a decreasing resistance with increasing pile displacement (see Figure 6.8.4). Similarly, a drilled shaft in
granular soil having a very strong soil-concrete bond may develop a soil motion which is similar to the pile motion
which again means that the soil is pulling away from the resistance force which then decreases rather early. Figure
6.8.5 shows a wave-up force curve which has a much lower magnitude after 2L/c than before the “valley”. These are
record characteristics which suggest that early unloading happens because of radiation damping.

One way of matching the behavior of Figure 6.8.4 and Figure 6.8.5 is by adding a dashpot underneath the regular
Smith model of static and dynamic resistance (BT for the toe and SK for the shaft). In effect, this model extension
allows the soil to move (see also Section 3.3.5). Intuitively, it is obvious that the static resistance can decrease (the
elasto-plastic spring can expand), even though the pile moves downward because the soil resistance forces push on
the dashpot and cause it to develop a velocity. An example of matching with this so-called radiation damping option
is discussed in the CAPWAP Help Section. Radiation damping matching, if improperly done or applied to the wrong
situation, may produce erroneous and dangerously non-conservative results. The shaft radiation damping model has
been correlated while for the toe no experience exists. It is reasonable, however, to use toe radiation damping (BT >
0.5) for end bearing in rock when damping factors required by a good match would become excessive (say higher
than 0.2 s/ft or 0.7 s/m).

Figure 6.8.4 Example of a load-set curve showing soil


resistance decrease with downward pile movement

CAPWAP Background Report Version 2014 70


Figure 6.8.5 Example of an upward traveling wave which displays
typical unloading characteristics

6.8.6 Plug mass at toe, PL


Soil trapped near the pile toe inside an open profile, or underneath a displacement pile, or any device of high mass
(toe plate, toe protection) attached to a pile toe exert an inertia force which is proportional to the pile bottom
acceleration. In the pile top force (or wave-up) record it therefore produces a positive force effect just before 2L/c and
a negative one just after 2L/c. Figure 6.8.6 shows an example where the plug effect (PL) can improve the match by
increasing the computed force just before, and decreasing it after, 2L/c. This effect makes the pile appear longer or
the wave speed slower. Other model parameters which can produce the same apparent delay of the toe reflected
impact wave include Pile Damping, PI, Soil Plug on Shaft, PS, reduction of the Overall Wave Speed (see Pile Model
PM), a series of Tension Slacks, and a Variable Time Increment.

Figure 6.8.6 Example of a match where a plug mass at the


toe, PL, could improve the match

6.8.7 Extra Pile Toes


Main characteristics of pile toe models include:
 No negative resistance
 Toe gap
 Quake independent of shaft quakes
 Damping independent of shaft damping
Extra pile toes therefore offer the analyst to model complex situations at the bottom the pile or at certain segments
along the shaft. The latter are generally assigned where the pile has a reduction of size such that soil displacements
occurs at some point along the pile. An example might be where an H pile protrudes out the bottom of a solid
concrete section; the soil immediately below the concrete will be moved by the impact and have an end bearing
effect. Sometimes the single toe resistance has to be modeled with at least two different toe resistance quakes to
produce a more rounded resistance response. It is then advisable to give the shaft resistance nearest the toe a value
consistent with those at the next 1 or 2 segments above (assuming the soil profile is relatively similar). The total end
bearing can then be split and the match improved by trying different quakes, toe resistance ratios, toe gaps and
damping options for each part of the combined toe model. The extra toe models have been further described in
Section 3.2.

CAPWAP Background Report Version 2014 71


6.9 Signal Matching with Non-Standard Pile Model Extensions
Standard pile model parameters include cross sectional area, elastic modulus, specific weight and, for output of unit
resistance values, the pile perimeter and the end bearing area (the latter two parameters do not affect the match;
they are only used for the calculation of unit resistance values). Elastic modulus and specific weight define the other
important standard pile parameter: the wave speed. During the matching process of a more complex situation (e.g., a
drilled shaft with an uncertain pile profile) it may become apparent that no reasonable match can be achieved with
soil model adjustments alone. In that case modifying one or more of the following pile model characteristics may lead
to an improved match.

• Added Impedance and/or Shaft Soil Plug;


• Compressive Slacks;
• Tensile Slacks;
• Residual Stress Analysis (RSA);
• Pile Damping (PI);
• Variable Time Increment.
The following discussion is intended to give some guidance as to when and why these parameters might be helpful.

6.9.1 Added Impedance and/or Shaft Soil Plug


Added impedance values are sometimes used to represent an enlargement in a drilled shaft while reduced
impedance may be a needed adjustment for piles with an unplanned cross sectional reduction or lower than average
product of wave speed and specific weight. These adjustments require that a certain soil resistance along the shaft is
fixed to values which are in reasonable agreement with the soil profile based static soil resistance calculations (within
some broad tolerance as we are never sure of true strength based solely on a boring). However, before doing so, the
user should be absolutely sure that matching without impedance increases or decreases or shaft soil plug
increases is not possible. Zn impedance increase will leave the travel time unaffected (and thus the 2L/c will occur at
the same time. Like the soil plug at the bottom, a shaft soil plug may cause a decrease in apparent wave speed or
apparent 2L/c time. Neither impedance change nor added soil shaft plug can be done to the pile top segment.

Examples where impedance increases are frequently employed are open ended pipe piles and H-piles. Drilled
shafts with uncertain cross sectional properties may require both impedance increases and decreases. Also, certain
splices or a follower to pile top connection may be most easily modeled with an impedance decrease. Another
example where impedance variations are needed are stabbing guides (used in offshore piling practice when adding
sections) or other short cross sectional increases near pile section connections, because using exact cross sectional
properties is not always successful because of shortcomings of the rather simple one dimensional pile model.
Impedance adjustments can lead quickly to a much improved match of such high frequency reflections. When using
impedance variations to model apparent cross sectional changes of a shaft with bulges or necks, improved
confidence in the solution may be achieved if the volume of concrete used is known. For that reason, the CAPWAP
software provides for a calculated volume (see also Section 2.6). In that case it may also be reasonable to calculate
the unit resistance values based on the apparent modified cross sectional values by converting the impedance
increase into the equivalent area increase in the pile profile PM with resulting changes to the perimeter and end
bearing values (accomplished in PM by selecting non-uniform on page one, and then selecting “Update Profile Using
Added Imp.” on page 2). Note that this option would then also modify the stiffness of the shaft in the static analysis.
Additionally, CAPWAP offers an automatic conversion of the original pile profile to the one that includes the modified
impedance. This should, however, not be done for impedance changes modeling a splice or added soil mass.

6.9.2 Compression and Tension Slacks


Slacks (see also Section 2.4) may be used where splices or cracks produce upward tension reflections of the impact
compression wave or they reflect downward the upward traveling tension wave. A slack prevents full transmission of
waves as long as the relative displacement of two neighboring segments is within the slack distance. Usually a slack
does not completely reflect a wave even though the slack is "open." In that case we speak of a slack with reduced
efficiency. Frequently, reinforcement ties two cracked sections together. In that case we can model a force limited
rather than displacement limited slack.

In Figure 6.9.1 the wave-up force shows before time 2L/c after impact a slight depression which suggests impedance
reductions, in this case due to tension cracks. The velocity record and the wave-up force show a "chopped off"
behavior at and after the time of impact plus 2L/c. This is indicative of an upwards traveling tension wave that was
partially downward reflected as it encountered cracks or a joint with limited tension capacity in the pile. Most likely,
the "chopped off" tension wave of Figure 6.9.1 is most easily matched with one or more tension slacks with limited
tension force transmission. Alternatively small tension slack deformations could be specified with small slack
efficiencies. A slack model which completely reflects a wave before it reaches its maximum slack extension, and

CAPWAP Background Report Version 2014 72


which after that completely transmits the wave, rarely leads to a successful match.

Figure 6.9.1 Record example of a pile with tension cracks

If a crack or slack is open at the beginning of the impact event, it will be closed by the compressive impact wave.
Although open pile cracks can be modeled by compressive slacks, an impedance reduction at one or more pile
segments is often more successful in matching the record, and considerably easier to model. Also pile damage which
may include loss of concrete or a bent steel pile, may be modeled by impedance changes. However, under certain
circumstances, a crack that is open when the compressive impact wave arrives, (which is then marked by a “beta” in
the PDA program) may have to be modeled with a compressive slack rather than by an impedance reduction. An
example of a measured wave-up that indicates a crack is shown in Figure 6.9.2. The crack appears to be open in the
beginning of the blow and therefore causes a slight tensile reflection before time 2L/c. The upward traveling, toe
reflected impact wave, being a tension wave, is also reflected when it reaches the crack which then reopens. The
2L/c reflection, in the upward traveling wave, therefore, has the characteristically flattened appearance. In this case
both compressive and tensile slacks (must be entered for neighboring segments) must be used to match the effect of
the crack or slack. The slacks may either be modeled as force limited or displacement limited.

Since it reflects the impact wave completely until the gap is closed, the compressive slack may generate a huge
reflection which can only be tempered with a reduced efficiency. Note that the CAPWAP model is limited in that it
does not allow for tension and compression slacks at the same segment. Instead neighboring segments should be
chosen for these two features. Fortunately, the match is usually not very sensitive to the location of the tension slack.
This is different for compressive slacks, where the location is well known from the initial reflection of the first
compression wave. In fact, it might be necessary to make smaller pile segments (increase the number of pile
segments) to produce a satisfactory match when a compressive slack must be modeled.

CAPWAP Background Report Version 2014 73


Figure 6.9.2 Record example of a potential compression slack plus
evidence of a tension slack at a splice location

6.9.3 Residual Stress Analysis, RSA


In general a CAPWAP solution using RSA provides a somewhat different resistance distribution but little change of
ultimate capacity prediction. For a long slender pile, RSA should always be tried with UN > 0.2. One benefit of RSA is
that its calculated residual pile stresses can be helpful in the interpretation of strain gage instrumentation of a
statically load tested pile. On the other hand RSA generally complicates the matching procedure substantially, Note
that CAPWAP’s RSA model may not give realistic answers for piles with slacks and that the toe gap is practically
ignored. RSA is more difficult to apply in cases of obvious large quakes.

6.9.4 Pile Damping, PI


Pile damping is generally fixed and is to be modified only if it can be explained by an unusually poor pile material
behavior (low compressive strength concrete or concrete contaminated by soil or not yet hardened). Under certain
circumstances increasing PI may produce a slowing of the wave similar to the effect of a series of tension slacks. In
that case it is much easier to work with PI. However, the user is urged to make sensitivity studies with other solutions
so as to provide for an improved confidence in an unusual solution. It should be appreciated that the wave speed WS
(or overall wave speed) must be properly determined rather than trying to adjust the 2L/c time by large changes in
PI..

6.9.5 Variable Time Increment and/or Wave Speed


This option, introduced in the 2014 version, was discussed in Section 2.8. For concrete piles which experience
micro-cracking or hairline tension cracking during driving or testing, the wave speed may be noticeably lower when
the stress wave becomes tensile. In effect this means that the overall wave speed and therefore the overall dynamic
elastic modulus is different after the first L/c time when tension develops in the pile.

Figure 6.9.3 and Figure 6.9.4 show two situations where obviously some impedance reduction is indicated in these
concrete piles (see arrows). While the relative impedance reduction in 6.9.3 is larger than in 6.9.4, both records show
that the velocity return wave has an appearance at 2L/c which is either that of a cut-off reflection or that of a lesser
slope than at impact. It is conceibable that more but smaller cracks would be hardly apparent in the first 2L/c record
portion, but would have an effect on the return wave at 2L/c (see also Figure 6.9.1).

CAPWAP Background Report Version 2014 74


Figure 6.9.3 Pile with strong impedance reduction

Figure 6.9.4 Pile with lesser impedance reduction

As we have seen earlier, tension slack modeling is one way to model the “chopped-off” appearaning return wave.
However the slowing down of the response and the reduced frequency of the pile-soil system due to an reduction in
apparent wave speed is often very difficult to match with tension slacks. Also, it is a lengthy process and the right
combination of segment number, slack size and slack efficiency values can only be determined in a tedious trial and
error procedure.

Alternatively, one could use a lower Overall Wave Speed in PM (corresponding to a reduced WC in the PDA).
However, that would change the wave speed already in the first part of the record when the pile is in compression
and behaves elastically. Furthermore, once stress levels decrease, the pile may again behave elastically and then the
original wave speed and thus the original time increment would be the correct choice. Ideally, one would choose the
elastic modulus (and therefore the wave speed) of each segment at the appropriate lower value when the segment is
in tension and at the elastic value when the segment is in compression. Unfortunately, with the rather rigidly elastic
characteristics pile model, such flexibility is not allowed. Instead one has to choose a lower wave speed for all pile
segments when the pile is more or less in tension and then switch all segments back to the higher wave speed when
the tension has ceased to exist.

The wave speed for any pile segment in CAPWAP Is represented by its segment length, ΔL, divided by time
increment, Δt. A lower wave speed is, therefore, represented by a longer time increment, since the length increment
does not change. Note, however, that since c 2 = E/ρ the elastic modulus can be calculated from E = ρ ΔL 2 / Δt2 . In
other words, E is inversely proportional to Δt2. Thus, increasing the time increment by a factor 2 would make the
equivalent elastic modulus 4 times lower than the elastic value. Such a large change cannot be attributed to the
concrete properties actually changing. Instead what is happening is that the wave is slowed down because it has to
open up the tension crack(s), and that takes time.

To help in cases where the wave speed changes after the intial impact event, the user can now specify a new time
increment to be used after a certain time. The variable time increment menu can be accessed by clicking on Input-
>Variable Time Increment. After entering the number of changes to be effected (for Example 2, one to turn on a
different increment and the second one to come back to the original one) and the appropriate timing and time
increment ratio, press APPLY to accept the change and perform a trial analysis. Please note that the variable time
increment affects the way the signals are displayed on the screen. Since this involves some complex scaling

CAPWAP Background Report Version 2014 75


calculations the screen display may not be current during automatic searches. However, it will be correct after the
search is finished.

There have been occasions when it appeared as though a time increment increase is also indicated for open ended
steel pipe piles which have a moving plug. For certain times the plug may move with the pile and by adding mass it
would cause a lower natural frequency of the system. In this case the starting time for the increased time increment
would probably be much later than 2L/c thereby only causing an improved match but not a significant change of
results.Also should the plug move again it may be desirable to turn of the changed time increment and, again, two or
more ime increment ratio inputs may be needed.

The variable time increment option is a very powerful tool for matching difficult situations. But because it is so
powerful, the user must use it wisely and only after all other possibilities have been checked and/or exhausted.
Ideally, a sensitivity study should be performed to be sure that the final results are reasonable and meaningful.

6.10 Additional Analysis Precautions


The final CAPWAP solution should conform to the following criteria. Discrepancies with these criteria are possible
but will often indicate an unrealistic solution.

(A) Smith damping factors are usually less than 0.4 s/ft (1.3 s/m). However, for records with very low peak
velocities or for very elastic, cohesive silts, plastic clays or silty clays, higher Smith damping factors may
be possible. In such soils to be conservative is more important than to adhere to rigid limits of parameters
which are reasonable for the average over all possible soil properties.
(B) In general damping factors less than 0.024 s/ft (0.08 s/m) should be avoided, because they could lead to
non-conservative results. However, when a very high end bearing occurs in a hard rock, low damping
factors may be reasonable.
(C) Blow count matching is generally recommended except where variable energies or changing capacities
make the observed blow count uncertain.
(D) Quakes should be greater than 0.04 inches (1 mm), primarily for numerical reasons. Shaft quakes, QS, are
generally less than 0.3 inches (7.5 mm) while there is no known limit on toe quakes. However, both shaft
and toe quakes have to be less than the associated pile displacement for each element or the predicted
capacity would not have been activated and therefore confirmed. Obviously, in these cases of limited pile
driving energy, the quakes are only a lower bound of their possible values (and the predicted capacity will
also be a lower bound value).
(E) Unloading quake multipliers are generally limited to a range between 0.3 and 1.0. The lower limit may
avoid numerical instability, the upper limit is, theoretically, a physical reality. However, exceptions are
possible. For example, for quakes larger than the standard 0.1 inches (2.5 mm), the unloading quake
multipliers may be chosen somewhat lower. For solutions with radiation damping, due to the quick load
shedding of the radiation damping model, occasionally CS values greater than 1 may be acceptable. In the
presence of toe gaps, which imply a higher toe quake than the the QT value itself, a CT > 1 is also
reasonable (but CT should be less than or equal to [QT + TG]/QT)
(F) As mentioned in (D), the predicted static capacity must be fully activated, i.e., the quakes must be less
than the pile displacements. Exceptions are first cycles of residual stress or multiple blow analyses.
However, the last cycles of these analyses must fully activate the predicted capacities.
(G) Shaft radiation damping analyses appear to be reasonable for cases of
o High blow counts (small sets per blow);
o Soils with silt or clay content;
o Very high damping factors due to a low wave-up after 2L/c ;
o Open ended piles with partial plugging behavior;
o Open ended pipes or H-piles with a potentially moving plug.
Toe radiation damping should be restricted to very hard soils or rock at the pile bottom. Radiation damping
tends to result in higher capacities than standard analyses and therefore has to be done with caution, and
the lower limit suggestion should be heeded to avoid non-conservative overpredictions.

(H) Residual stress analysis is only meaningful if the UN value is at least 0.2. In contrast to wave equation
experience, CAPWAP with RSA normally does not calculate higher capacities than standard analyses.
However, the resulting resistance distribution and quake values may be quite different.

(I) While the goal is to obtain the best possible match quality, the CAPWAP result is only as good as its soil
model. The CAPWAP results, therefore, must be checked by comparing them with other information
available on the prevailing soil type and local experience. In that context, if the match quality is insensitive to
variations in capacity, then conservative solutions should be chosen in cohesive soils and/or when blow
counts are low and more aggressive solutions in dense granular soils and/or very high blow counts.

CAPWAP Background Report Version 2014 76


7 CAPWAP LIMITATIONS AND APPLICATION RECOMMENDATIONS

7.1 CAPWAP (Dynamic Pile Testing) Limitations


CAPWAP is widely considered the best means for analyzing data from dynamic pile testing both for bearing capacity
and pile integrity determination. The following description of limitations of bearing capacity determination and
potential mitigations are therefore those of dynamic pile testing in general. For example, we know that dynamic pile
testing can be used to obtain the soil resistance at the time of testing and, therefore, the CAPWAP predicted pile
bearing capacity is also representative for the forces existing in the soil-pile interface at the time of testing. If therefore
the intent is to determine the long term pile capacity then it is necessary to analyze dynamic data that has been taken
during a restrike test that has been conducted a sufficiently long time after pile installation such that the effects of soil
setup and relaxation are automatically included in the test results. These and other possible error sources are
discussed in the following in no particular order.

7.1.1 Underprediction Due to Lack of Hammer Energy


7.1.1.1 Mitigation by the superposition of EOD and BOR method
Underprediction is a common problem if the test was conducted with insufficient energy. This condition is generally
recognized by a high blow count (>120 blows/ft) or small set per blow (<2.5 mm). Unfortunately, it is not possible to
give a recommendation as to how much the underprediction would be; in fact sometimes the CAPWAP correlations
are very good even for low energy blows, particularly in cohesive soils. It is, therefore, not advisable to rely on higher
capacities than predicted just because the set per blow during the test was low. However, to be sure it has become
fairly common to restrike test a pile with a hammer that is more powerful than the installation hammer. If such a more
powerful hammer is not available then it may be necessary to utilize the superposition method (Hussein et al., 2002)
for a correction of the capacity prediction. This method is based on the assumption that the end bearing was fully
activated at the end of driving while the restrike activated the shaft resistance, but not all of the end bearing. If that is
so, then adding the end-of-drive end bearing to the early restrike shaft resistance would lead to a more realistic total
capacity. Obviously the analyst must be very careful in determining the resistance distribution of both end of drive
and early restrike records which is easier if the toe has a large quake at end of driving (at restrike we cannot expect
that because of the partial activation). Alternatively, for the end of drive toe resistance, the shaft resistance at the
bottom segment should be similar to the shaft resistance above those segments for conservatism.

7.1.1.2 Underprediction mitigation by the superposition of BOR and EOR method


It has been pointed out earlier that a potential for underprediction exists when the energy increases blow by blow
while the soil resistance decreases during the early blows of a restrike test. In that case the hammer may be powerful
enough for activating all resistance, except that it is a slow starter. Thus, the early blows probably only fully activate
the resistance in the upper soil layers. Later impacts would be able to activate all of the resistance, but then soil setup
has been lost in the upper soil layers. Underprediction would be the result if either an early or a late record is
analyzed. Superposition can help in this case too by combining the results from early and late restrike blows. This is
best done by transferring the final resistance results of all records analyzed to a spread sheet. Figure 7.1.1 shows a
made-up example of how the superposition leads to a resistance envelope which is composed of the largest values of
the individual segment capacity values (the legend also indicates the total capacity prediction for each blow).

CAPWAP Background Report Version 2014 77


Resistance force in k
0 5 10 15 20 25 30
0

2 BLow 1 - 51 k

Blow 2 - 59 k
4
Blow 4 - 60k
Depth

6 Last Blow - 67 k

Envelope - 90 k
8

10

12

Figure 7.1.1: Synthetic example of superposition of several restrike results


Figure 7.1.1 Synthetic example of superposition of several restrike results

7.1.1.3 Precautions for the use of superposition


Caution: To avoid overprediction, it is very important that any type of superposition is only done in cases where
the restrike blow count is at refusal (say greater than 120 blows/ft) or the set per restrike blow is very small (say
less than 2.5 mm/blow).

Also the EOD end bearing plus BOR shaft resistance superposition method should only be used
(a) For cases where the pile toe is deeply embedded in granular soils (for cohesive soils the end bearing should
be insignificant)
(b) For cases with pile toes supported by non-relaxing soils (e.g., not in saturated very dense fine grained soils
which tend dilate and thus develop negative pore water pressures and temporarily high effective stresses or
in certain shales which tend to relax). Experience with pile driving in identical soils, therefore, is necessary!

Additionally, BOR and EOR shaft resistance superposition should only be used when the resistance distribution
along the shaft has been very carefully evaluated. For if the high values of high resistance layers were calculated
for different segments for different blows then severe overprediction will result.

7.1.2 Underprediction Due to Lack of Setup for EOD or Early BOR Tests
In many or most instances soil setup occurs along the pile shaft after the driving of the pile has ended. Reasons for
setup can be simply a reduction of excess pore pressures and that may happen within a relatively short time in
sands. Thus a restrike test after a few hours may indicate a much higher capacity than the end of driving. However, in
fine grained soils the setup phenomenon may take a much longer time, under certain circumstances one month or
more. In fact it is generally expected that soil setup in fine grained soils occurs linearly with the logarithm of time.

While easier said than done this very frequent source of underprediction can be avoided (a) if capacity is calculate
from restrike tests (rather than from EOD) and (b) if the restrike tests are performed a sufficient time after pile driving.
Obviously, when comparing static and dynamic load tests then this only makes sense if the tests were done at
comparable times (Likins et al., 1996).

7.1.3 Underprediction due to Plugging in Sands


As has been mentioned earlier in this report, open end pile profiles (pipes, H-piles) may behave differently during
static and dynamic load applications: In cohesionless soils or other soils with relatively with high end bearing, the soil
plug inside the pile (Note: a soil plug moves at least to some degree with the pile; if it does not move with the pile it is

CAPWAP Background Report Version 2014 78


called a soil column) may slip during the dynamic test due to the high inertia of the soil plug plus the end bearing
forces and, therefore, full end bearing will not develop during driving, particularly for larger diameter profiles where
the soil inertia can be very high. However, during static loading, friction inside the pipe or between the flanges of an H
pile may prevent the soil plug from slipping and, thus, be able to support full end bearing. In general this means that
the dynamically determined capacity is less than the statically determined one. Because the slipping of the soil plug is
primarily the result of high inertia the underprediction may be more severe for larger piles than for smaller piles; in
fact there is consensus that a pile well embedded in a dense layer will remain plugged during the dynamic load
application if its diameter is 30 inches (750 mm) or less. For very large piles, say more than 96 inches (2500 mm)
diameter, plugging may not even happen in the static case and may be unreliable for intermediate pile sizes.

It is not easy to avoid this error source and research is ongoing. However, it can be tried to restrike with low stroke,
large mass hammers which impose low accelerations. If static test results indicate higher capacities then it may be
possible to rely on static analyses in conjunction with the dynamic tests to assess capacity. Recommendations as to
how to do this are beyond the scope of this report.

7.1.4 Overprediction Due to High Energies


While underprediction of capacity is conservative, overprediction could possibly cause distress in the structure
supported by piles with insufficient capacity and is, therefore, a much more dangerous problem. Fortunately it occurs
less frequently than underprediction. Overprediction occasionally occurs if the hammer energy during the test is so
high that it produces an equivalent blow count less than 24 blows/ft or a set per blow greater than 12 mm. The reason
is the possible misinterpretation of high damping components as static resistance. For that reason is it recommended
to reduce the restrike energy if possible when the hammer is too powerful.

7.1.5 Overprediction Due to End Bearing Relaxation


Relaxation has been mentioned; in this general case, analysis of early restrike blows shows that the capacity is lower
than at the end of driving or the end of restrike testing. Therefore, performing restrike tests with uniform hammer
energies is the best means of assuring accurate predictions. Morgano and White, 2004, described this phenomenon
and its mitigation.

7.1.6 Overprediction in Highly Plastic Soils


In general, CAPWAP does a very good job at separating damping resistance from static resistance effects. However,
in certain highly cohesive or plastic clays or silty clays (it is generally not possible to identify these problem soil types
from typical soil profiles) rate effects can still lead to an overprediction which is typically limited to about 30 or 40%.
To avoid this error, check the following
(a) Does experience exists with dynamic testing and static test results.
(b) Is the calculated end bearing reasonable (it should not be very high); if not reduce it.
Where no experience and the calculated end bearing is reasonable or has been reduced, either a static load test
should be performed or the calculated capacity should be reduced by 30%.

Anecdotal evidence also exists that cohesive soils which do not show a clearly lower shaft resistance during driving
than during restrike (i.e., those that do not exhibit soil setup) tend to have higher rate dependencies than those that
clearly develop setup.

7.1.7 Overprediction Due to Soil Column Adhesion in Cohesive Soils


In 7.1.3 we discussed the underprediction potential due to a slipping plug in large open end pipe piles and
cohesionless soils. In cohesive soils occasionally the opposite happen. During the dynamic event adhesion inside the
pipe may cause a high driving resistance, however, when statically loaded the soil column will not encounter much
end bearing and therefore less resistance than encountered during the dynamic event. Thus the dynamic resistance
is in this case greater than the static one and this effect may be aggravated in particularly plastic clays or silty clays
by rate effects.

7.1.8 Underprediction of Settlements


Like the total capacity prediction and corresponding shaft resistance and end bearing components, CAPWAP
calculated load-displacement curves correspond to the soil condition at the time of testing. It must be recognized that
the dynamic test occurs in a short time period. Thus, settlement predictions based on CAPWAP results (e.g., using
the CAPWAP calculated load-displacement curve) should consider additional long-term effects such as consolidation
or creep of the soils surrounding the pile or potentially compressive layers some distance below the pile toes, as well
as group effects of a cluster of piles in a pile cap.

CAPWAP Background Report Version 2014 79


7.1.9 Errors in Calculated Resistance Distribution and Uplift Capacity
The separation of shaft from toe resistance is difficult near the pile toe. Even direct strain measurements on the pile
cannot determine the correct end bearing during static or dynamic tests. Particular care must, therefore, be exercised
when making uplift predictions from CAPWAP calculated shaft resistance. In general it is wise to use only 80% of the
calculated shaft resistance after reducing it by unusually high values that may have been calculated for one or two
shaft segments at and above the toe. To avoid unrealistic expectations of compressive shaft resistance and/or uplift
capacity it is therefore strongly recommended to check the predicted unit resistance values and compare them with
commonly recommended limits, and to conduct a parameter study of the optimum ratio of shaft resistance to end
bearing. Short piles may not fail in shear during uplift but rather due to lack of mass in the soil cone surrounding the
pile.

7.1.10 The Uniqueness Question


CAPWAP is an inverse analysis and for this reason different solutions can be proposed. Generally, the correct
capacity is calculated for the best match quality value and the same capacity result is achieved for the same match
quality value. Obviously, different capacity values would be obtained if the best match has not been achieved.
However, the sensitivity of the capacity value to the small changes in MQ (match quality) is generally not very strong
and preliminary results from the fully automated iCAP method have demonstrated this point.

Section 7.1.9 alluded to the fact that neighboring segment resistance values can be indeed be varied without causing
much of a change of the match quality. It is, therefore, strongly recommended to perform sensitivity analyses and
select the one result that corresponds most closely to what can be reasonably expected for a certain soil profile.

7.2 Comparing Apples with Oranges


Capacity and resistance distribution comparisons from different test types (static, rapid, dynamic) are always
challenging. The tests have to be done at a comparable time and on the same pile. If different piles are subjected to
comparison tests then differences must be expected, particularly when the piles are drilled, or not driven to the same
blow counts. However, when the same pile is tested then the first test is on an unloaded pile while the following test is
on a preloaded pile (and thus follows the reload stiffness).

Measurements of strains in a test pile during a static test are quite reliable on steel piles; on concrete piles, however,
the elastic modulus is often not well known and forces calculated from strain measurements, therefore, are potentially
inaccurate. This problem is aggravated for cast-in situ piles which are frequently of variable cross section.
Furthermore, experience has shown that the accuracy of static testing, if improperly performed, can be severely
compromised. For example, installation of reaction piles after the installation of the load test pile (particularly when
installed with a vibratory hammer) or short reaction piles or reaction piles installed too close to the test pile can
strongly affect the measured static capacity. Hydraulic pressure gages tend to indicate excessively high forces
because of jack friction and electronic load cells tend to drift with uncertain errors. Thus, it is always wise to critically
review all static testing results.

7.3 Additional Design Considerations


Calculation of bearing capacity is only one aspect of deep foundation design. Obviously, settlement calculations are
important and as pointed out above, have to consider the possibility of long term effects such as creep or
consolidation.

Numerous other factors also have to be considered in pile foundation design. Some of these considerations include

1. additional pile loading from downdrag or negative skin friction, lateral and uplift loading requirements
2. effective stress changes (due to changes in water table including artesian pressures, excavations, fills or other
changes in overburden),
3. long term settlements in general and settlement from underlying weaker layers and/or pile group effects,
4. loss of shaft resistance due to scour or other effects,
5. loss of structural pile strength due to additional bending loads, buckling (the dynamic loads general do not cause
buckling even though they may exceed the buckling strength of the pile section), corrosion etc.

7.4 Factors of Safety


Run to failure, static or dynamic load tests yield an ultimate pile bearing capacity, R u. If this failure load were applied
to the pile during its service, then excessive settlements would occur. Therefore, it is absolutely necessary that the
actually applied load (also called the design load, R d, or working load or safe load) is less than R u. In most soils, to
limit settlements, it is usually required that Ru exceeds Rd by at least 50% Thus

CAPWAP Background Report Version 2014 80


Ru ≥ 1.5 Rd,
or the Factor of Safety has to be at least 1.5.

Unfortunately, neither applied loads nor Ru are exactly known. One static load test may be performed at a site, but
that would not guarantee that all other piles have the same capacity and it is to be expected that a certain percentage
of the production piles have lower capacities, either due to soil variability or due to pile damage. If, for example,
dynamic pile tests are performed on piles in shale only a short time after pile installation, then, because of relaxation,
the test capacity may be higher than the long term capacity of the pile.

Not only bearing capacity values of all piles are unknown, even loads vary considerably and occasional overloads
must be expected. We would not want a structure to become unserviceable or useless because of either an
occasional overload or a few piles with low capacity. For this reason, and to avoid being overly conservative, which
would mean excessive cost, modern safety concepts suggest that the overall factor of safety should reflect both the
uncertainty in loads and resistance. Thus, if all piles were tested statically and if we carefully controlled the loads, we
might be safe with F.S. = 1.5. However, in general, depending on the building type or load combinations and as a
function of quality assurance of pile foundations, a variety and in general, for settlement reasons higher, Factors of
Safety have been proposed.

For example, years ago, for highway related projects the American Association of State Highway and Transportation
Officials (AASHTO) and the USA Federal Highway Administration had proposed (the proposal is still reasonable even
though since 2007 AASHTO switched to an Load and Resistance Factor (LRFD) approach) the following factors of
safety:

F.S. = 1.90 with both static and dynamic load tests


F.S. = 2.00 for static load test with wave equation;
F.S. = 2.25 for dynamic testing with signal matching;
F.S. = 2.50 for indicator piles with wave equation;
F.S. = 2.75 for wave equation analysis;
F.S. = 3.50 for Gates or other dynamic formula.

Based on these recommendations, in the absence of a static load test (or in absence of static load test experience in
a comparable geology) a factor of safety of 2.25 is reasonable if dynamic testing with CAPWAP is done on enough
piles to assure that site variability is not an issue. Note that these guidelines do not specify a percentage of piles to
test. Some practitioners have suggested 5% while some might suggest only 2% of all piles be tested dynamically.
Since frequently the piles are responsible for the main cost of the foundation, more testing may result in optimizing
the foundation and actually reduce the overall cost. Tests should be specified to explore the variability of the soils
across the site as well as to observe the consistency of the hammer performance with time for large projects where
installation of the piles takes weeks to accomplish. It is interesting to note that in 2007 AASHTO switched to an
LRFD requirement and required for each site condition dynamic testing of a minimum of “2% or 2 piles, whichever is
larger”. Further, AASHTO also equated testing of 100% of the piles dynamically with performing a static test (the
equivalent in allowable stress design would be to use the same safety factor of 2.00 for either one or only a few static
tests or dynamic tests of 100% of the piles.

It is the foundation designer’s responsibility to identify design loads together with the adopted safety factor concept
and the associated construction control procedure. The required factors of safety should be included in design
drawings or specifications together with the required testing, or better still show the required ultimate capacity
(ultimate capacity is called “nominal resistance” in AASHTO’s LRFD guide line specification and “characteristic
resistance” in the European Code, EC7). Only then can the contractors bidding for the work develop the most
economical solution and develop a program that is sensitive to the economic impact of increased testing for lower
required pile capacities. This will also help to reduce the confusion that often exists on construction sites as to design
loads and required capacities. In any event, it cannot be expected that the test engineer is aware of and responsible
for the variety of design considerations that must be met to find the appropriate factor of safety.

CAPWAP Background Report Version 2014 81


APPENDIX A: REFERENCES

Chin, V.B.L., and Seidel, J.P., (2004), “An Experimental Study into the Viscous Damping Response of Pile-Clay
Interfaces”, Proceedings of the 7th Int. Conf. on the Appl. of Stress-wave Theory to Piles 2004, p. 435-446.
Coyle, H.M., and Gibson, G.C., (1970), "Empirical Damping Constants for Sands and Clays," ASCE Journal of Soil
Mechanics and Foundation Division.
Hussein, M.H., Sharp, M., and Knight, W.,F., 2002. The Use of Superposition for Evaluating Pile Capacity. Deep
Foundations 2002, An International Perspective on Theory, Design, Construction and Performance; GSP No. 116,
O’Neill and Townsend Eds., ASCE, Orlando, FL; 6-21.
Likins, G. E., Rausche, F., Thendean, G., Svinkin, M., September 1996. CAPWAP Correlation Studies. Proceedings
of the 5th Int. Conf. on the Application of Stress-wave Theory to Piles 1996: Orlando, FL; 447-464.
Likins, G. E., Liang, L., Hyatt, T., September 2012. Development of Automatic Signal Matching Procedure - iCAP®.
Proceedings from Testing and Design Methods for Deep Foundations; IS-Kanazawa: Kanazawa, Japan; 97-104.
Likins, G. E., Rausche, F., 2004. Correlation of CAPWAP with Static Load Tests. Proc. of the 7 th Int. Conf. on the
Application of Stress-wave Theory to Piles. Petaling Jaya, Selangor, Malaysia, Keynote Lecture, 153-165.
Likins, G. E., Rausche, F., DiMaggio, J., Teferra, W., September, 1992. “A solution for high damping constants in
sands”, 4th Int. Conf. on the Application of Stress-wave Theory to Piles, The Netherlands, p. 117-120.
Morgano, C.M. and White, B., 2004. Identifying soil relaxation from dynamic testing. Proc. of the 7 th Int. Conf. on the
Application of Stress-wave Theory to Piles. Petaling Jaya, Selangor, Malaysia, 415-421.
Rausche, F., 1970. Soil Resistance Predictions from Pile Dynamics, Ph.D., Thesis, Case Western Reserve
University, Cleveland, Ohio.
Rausche, F. Moses, F., and Goble, G., 1972. Soil Resistance Predictions from Pile Dynamics. Journal of the Soil
Mechanics and Foundations Division, ASCE, Reprinted in Current Practices and Future Trends in Deep Foundations,
Geotechnical Special Publication No. 125, DiMaggio, J. and Hussein, M., eds., 418 – 440.
Rausche, F., January 1983. CAPWAP Analysis Using the Characteristics Approach. PDA User's Day: Philadelphia,
PA; 1-7.
Rausche, F., Likins, G.E., and Goble, G.G., (1994), "A Rational and Usable Wave Equation Soil Model Based on
Field Test Correlation", Proceedings, Design and Construction of Deep Foundations, Federal Highway
Administration, Washington, D.C.
Rausche, F., Likins, G. E., Liang, L., Hussein, M.H., January 2010. Static and Dynamic Models for CAPWAP Signal
Matching. The Art of Foundation Engineering Practice, Geotechnical Special Publication No. 198, Hussein, M. H., J.
B. Anderson, W. M. Camp, eds., American Society of Civil Engineers: Reston, VA; 534-553.
Rausche, F., Nagy, M., Likins, G. E., September 2008. Mastering the Art of Pile Testing. 8th Int. Conf. on the
Application of Stress-wave Theory to Piles 2008: Lisbon, Portugal; 19-32.
Rausche, F., Nagy, M., Webster, S., Liang, L., May 2009. CAPWAP and Refined Wave Equation Analyses for
Driveability Predictions and Capacity Assessment of Offshore Pile Installations. Proceedings of the ASME 28 th
International Conference on Ocean, Offshore and Arctic Engineering: Honolulu, Hawaii; 1-9.
Rausche, F., Robinson, B., and Liang, L., 2000. Automatic Signal Matching with CAPWAP. 6 th Int. Conf. on the
Application of Stress-wave Theory to Piles, Sao Paulo, Brazil, 53-58.
Rausche, F., Pezzetti, G., Jannaci, C., and Klesney, A., 2006. Dynamic Analysis of Follower Driven Piles for the
Venice Flood Gate Project. 10th Int. Conf. on Piling and Deep Foundations, The Deep Foundations Institute,
Amsterdam, Netherlands, 763-772.
Smith, E.A.L., (1950), "Pile Driving Impact," Proceedings, Industrial Computation Seminar, September 1950,
International Business Machines Corp., New York, N.Y., p. 44.
Smith, E.A.L., (1960), “Pile Driving Analysis by the Wave Equation,” Journal of the Soil Mechanics and Foundations
Division, ASCE, Volume 86.

CAPWAP Background Report Version 2014 82

You might also like