Skip to main content

Full text of "DTIC AD1014538: Simplified Dynamic Structural Time History Response Analysis of Flexible Approach Walls Founded on Clustered Pile Groups Using Impact_Deck"

See other formats


Information and Technology Laboratory erdc/itltr- 16 -i 



US Army Corps 
of Engineers® 

Engineer Research and 
Development Center 



INNOVATIVE SOLUTIONS 
for a safer, better world 


Navigation Systems Research Program 

Simplified Dynamic Structural Time-History 
Response Analysis of Flexible Approach Walls 
Founded on Clustered Pile Groups Using 
Impact.Deck 

Barry C. White, Jose Ramon Arroyo, and Robert M. Ebeling July 2016 


Cross- 

Section 




Barge A 
Impact 


Cross-Section 


McAlpine Alternative 
Impact Flexible Wall 


O 


— Q 0 - <g r( p - Q ( p — 


Barge 

Impact 


Approved for public release; distribution is unlimited. 
































































The US Army Engineer Research and Development Center (ERDC) solves the 
nation’s toughest engineering and environmental challenges. ERDC develops innovative 
solutions in civil and military engineering, geospatial sciences, water resources, and 
environmental sciences for the Army, the Department of Defense, civilian agencies, and 
our nation’s public good. Find out more at www.erdc.usace.army.mil . 

To search for other technical reports published by ERDC, visit the ERDC online library 
at http://acwc.sdp.sirsi.net/client/default . 




Navigation Systems Research Program 


ERDC/ITL TR-16-1 
July 2016 


Simplified Dynamic Structural Time-History 
Response Analysis of Flexible Approach Walls 
Founded on Clustered Pile Groups Using 
Impact.Deck 

Barry C. White and Robert M. Ebeling 

Information Technology Laboratory 

U.S. Army Engineer Research and Development Center 

3909 Halls Ferry Road 

Vicksburg, MS 39180-6199 

Jose Ramon Arroyo 

Department of General Engineering 
University of Puerto Rico 
Mayaguez, PR 00681 


Final report 

Approved for public release; distribution is unlimited. 


Prepared for U.S. Army Corps of Engineers 
441 G. Street NW 
Washington, DC 20314-1000 

under Work Unit Number 448769 



ERDC/ITL TR-16-1 


ii 


Abstract 

Flexible approach walls are being considered for retrofits, replacements, or 
upgrades to Corps lock structures that have exceeded their economic 
lifetime. This report discusses a new engineering software tool to be used 
in the design or evaluation of flexible approach walls founded on clustered 
pile groups and subjected to barge train impact events. 

This software tool, Impact_Deck, is used to perform a dynamic, time- 
domain analysis of three different types of pile-founded flexible approach 
walls: an impact deck, an alternative flexible approach wall, and a guard 
wall. Dynamic loading is performed using Impact_Force time histories 
(Ebeling et al. 2010). This report covers the numerical methods used to 
create this tool, a discussion of the graphical user interface for the tool, 
and an analysis of results for the three wall systems. 

The results of analyzing the three wall systems reveals that dynamic 
evaluations should be performed for these structures because of inertial 
effects occurring in the wall superstructure and substructure. These 
inertial effects can cause overall and individual response forces that are 
greater than the peak force from the Impact_Force time history. 

This report also discusses the advantages of load sharing between multiple 
pile groups in an approach wall substructure. In the case of Lock and Dam 
3, the peak reaction force for any individual pile group was 11% of the peak 
impact load. 


DISCLAIMER: The contents of this report are not to be used for advertising, publication, or promotional purposes. 
Citation of trade names does not constitute an official endorsement or approval of the use of such commercial products. 
All product names and trademarks cited are the property of their respective owners. The findings of this report are not to 
be construed as an official Department of the Army position unless so designated by other authorized documents. 

DESTROY THIS REPORT WHEN NO LONGER NEEDED. DO NOT RETURN IT TO THE ORIGINATOR. 




ERDC/ITL TR-16-1 iii 


Contents 

Abstract.ii 

Figures and Tables.vi 

Preface.xiii 

Unit Conversion Factors.xv 

1 Dynamic Structural Time-History Response Analysis of a Flexible Approach Wall 

Supported by Clustered Pile Groups during Impact with a Barge Train.1 

1.1 Introduction - Glancing Impact Blows and Flexible Approach Wall Structural 

Systems.1 

1.2 Examples of the Next Generation Flexible Approach Walls.1 

1.2.1 Lock and Dam 3 .1 

1.2.2 McAlpine flexible wall .3 

1.2.3 Guard walls .4 

1.3 Overview of Dynamic Time-Flistory Response Analysis of a Flexible Beam 

Supported Over Elastic-Plastic Spring Supports.4 

1.4 Report Contents.8 

2 Impact Deck Approach Wall - Lock and Dam 3 Example.11 

2.1 Introduction.11 

2.2 Lock and Dam 3 - Physical Model.11 

2.3 Lock and Dam 3 - Construction Drawings.18 

2.4 Lock and Dam 3 - Mathematical Model.18 

2.5 Nonlinear force-deflection relationship for the spring supports.31 

2.6 Solving for the motion of the structure.32 

2.7 Validation of lmpact_Deck Computer Program.32 

2.8 Numerical Example of the Elastic-Plastic Response Using lmpact_Deck.34 

2.9 lmpact_Deck GUI results.36 

2.10 Final Remarks.56 

3 An Approach Wall with Impact Beams on Nontraditional Pile Supported Bents - 

McAlpine Example.57 

3.1 Introduction.57 

3.2 Alternative Flexible Approach Wall - Physical Model.57 

3.3 McAlpine Alternative Flexible Approach Wall - Mathematical Model.58 

3.4 Nonlinear force-deflection relationship for the springs supports.60 

3.5 Solving for the motion of the structure.62 

3.6 Validation of lmpact_Deck Computer Program.63 

3.7 Numerical Example of the Elastic-Plastic Response Using lmpact_Deck.66 

3.8 lmpact_Deck GUI results.67 

3.9 Final Remarks.84 

4 Traditional Impact Beam Guard Walls.85 





































ERDC/ITL TR-16-1 iv 


4.1 Introduction.85 

4.2 Guard Walls - Physical Model.85 

4.3 Guard wall - Mathematical Model.86 

4.4 Nonlinear force-deflection relationship for the springs supports.88 

4.5 Solving for the motion of the structure.90 

4.6 Validation of lmpact_Deck Computer Program.90 

4.7 Numerical Example of the Elastic-Plastic Nonlinear Response Using 

lmpact_Deck.92 

4.8 lmpact_Deck GUI results.93 

4.9 Final Remarks.110 

5 lmpact_Deck Graphical User Interface (GUI).Ill 

5.1 Introduction.Ill 

5.2 Geometry Tab.112 

5.3 Impact Time History Tab.118 

5.4 Beam Properties Tab.119 

5.5 Pile Cluster Spring Tab.121 

5.6 Analyze Tab.123 

5.7 Output Tab.126 

5. 7.1 FEO Nodal Output . 127 

5.7.2 FEO Element Output . 133 

5.7.3 FEO Pile Group Response . 138 

5.7.4 Run Information . 143 

5.8 Example: Geometry Input for the Impact Deck at Lock and Dam 3.144 

5.9 Final Remarks.148 

6 Conclusions and Recommendations.149 

References.153 

Appendix A: Lock and Dam 3 - Equations of Motion for the Mathematical Model.155 

Appendix B: McAlpine Alternative Flexible Wall - Equations of Motion for the 

Mathematical Model.165 

Appendix C: Guard wall - Equations of Motion for the Mathematical Model.179 

Appendix D Push-over analysis for batter-pile bent system.189 

Appendix E: Formulation for the rotational spring stiffness for the McAlpine flexible 

approach wall clustered group of vertical piles model.224 

Appendix F: HHT-a method.232 

Appendix G: Wilson-6 Method.235 

Appendix H: Member End Release Details for Load Applied at the End Release Node.238 


Appendix I: Rayleigh Damping 


240 





































ERDC/ITL TR-16-1 


v 


Appendix J: Key lmpact_Deck Program Variables.245 

Report Documentation Page 




ERDC/ITL TR-16-1 vi 


Figures and Tables 

Figures 

Figure 1.1 Front-end cells of Lock and Dam 3.2 

Figure 1.2 Lock and Dam 3 cross-section and plan view.3 

Figure 1.3 McAlpine flexible walls.4 

Figure 1.4 Guard wall schematic drawing.5 

Figure 1.5 Barge train impacting at a fixed impact position along the simply supported, 
flexible impact beam mathematical model with the barge train oriented at an approach 
angle 0 to the wall’s XGiobai axis (plan view).6 

Figure 1.6 Example of an impact pulse-force time history.6 

Figure 1.7 Barge impact point force moving along the wall from initial contact time ti to 

final contact time t 2 ..7 

Figure 2.1 Front end cell of Lock and Dam 3.12 

Figure 2.2 Pinned connection between the circular cell and the Impact Deck at Lock and 

Dam 3.13 

Figure 2.3 Arrangement of piles at Lock and Dam 3.13 

Figure 2.4 Impact Deck supported over piles at Lock and Dam 3.14 

Figure 2.5 An upstream view starting at the concrete cell of the pile-founded flexible 

approach wall at Lock and Dam 3.14 

Figure 2.6 Precast bases before installation.15 

Figure 2.7 Precast bases during installation.15 

Figure 2.8 Precast bases connected to piles.16 

Figure 2.9 Construction joint between concrete block segments (axial, shear and moment 
transfer connection).16 

Figure 2.10 Construction joints.17 

Figure 2.11 Massive concrete circular cell.17 

Figure 2.12 Pile-founded flexible impact deck structure.18 

Figure 2.13 Lock and Dam 3 guide wall plan view 1 of 2.19 

Figure 2.14 Lock and Dam 3 guide wall plan view 2 of 2.20 

Figure 2.15 Lock and Dam 3 guide wall detail of pile layout and end cell.21 

Figure 2.16 Lock and Dam 3 guide wall section view.22 

Figure 2.17 Lock and Dam 3 guide wall plan - 5.23 

Figure 2.18 Lock and Dam 3 guide wall plan - 6.24 

Figure 2.19 Lock and Dam 3 flexible approach wall.25 

Figure 2.20 (a) Typical 3-D segment of the impact-deck beam element, (b) Impact force 
applied to the Impact Deck, (c) Typical 3-D beam element.25 

Figure 2.21 Typical 2-D beam element used in lmpact_Deck.26 

Figure 2.22 Lock and Dam 3 mathematical model.27 

Figure 2.23 First beam element used in Lock and Dam 3 mathematical model.27 

Figure 2.24 Last beam element used in Lock and Dam 3 mathematical model.27 


































ERDC/ITL TR-16-1 vii 


Figure 2.25 Description of two beam elements connected at the inter-monolith 

connection.28 

Figure 2.26 Beam elements numbering in a typical internal monolith.28 

Figure 2.27 Cross-section of Lock and Dam 3.29 

Figure 2.28 Transverse and longitudinal push-over results for a single row of three piles 
aligned in the transverse direction.30 

Figure 2.29 Force-displacement relation of the spring support.32 

Figure 2.30 Force time history of Winfield Test # 10.33 

Figure 2.31 Validation of lmpact_Deck against SAP2000.34 

Figure 2.32 Dynamic response of the transverse spring located at x = 402.896 ft. .35 

Figure 2.33 Dynamic response of the transverse spring located at x = 402.896 ft. .36 

Figure 2.34 Reprint of the Figure 4-4 transverse direction of loading push-over analyses 

from Ebeling et al. (2012); fixed head results in green and pinned head results in blue.37 

Figure 2.35 lmpact_Deck GUI Table of maximum nodal displacements for the L&D3 

example impact deck.39 

Figure 2.36 Transverse nodal displacement time histories for node 86.40 

Figure 2.37 Longitudinal nodal displacement time histories for node 137.40 

Figure 2.38 Rotational nodal displacement time histories for node 93.41 

Figure 2.39 Transverse wall displacements at 0.26 sec.42 

Figure 2.40 Longitudinal wall displacements at 0.21 sec.42 

Figure 2.41 Rotational wall displacements at 0.26 sec.43 

Figure 2.42 lmpact_Deck GUI table of element minimum and maximum axial forces for 

the L&D3 example impact deck.43 

Figure 2.43 lmpact_Deck GUI table of element minimum and maximum shear forces for 

the L&D3 example impact deck.44 

Figure 2.44 lmpact_Deck GUI table of element minimum and maximum moments for the 
L&D3 example impact deck.44 

Figure 2.45 Axial-force time histories for element 82.45 

Figure 2.46 Axial-force time histories for element 81.46 

Figure 2.47 Shear-force time histories for element 82.46 

Figure 2.48 Shear-force time histories for element 81.47 

Figure 2.49 Moment time histories for element 81.47 

Figure 2.50 Moment time histories for element 82.48 

Figure 2.51 Wall axial forces at 0.2 sec.48 

Figure 2.52 Wall shear forces at 0.2 sec.49 

Figure 2.53 Wall moments at 0.18 sec.49 

Figure 2.54 Table of pile group response maximum forces and moments and their time.50 

Figure 2.55 Table of pile group response maximum displacements and their time.51 

Figure 2.56 Table of pile group responses for each pile group individually and summed 

(not shown) at time 0.26 sec.51 

Figure 2.57 Transverse pile group responses for the pile group at node 85 and at time 

0.24 sec.52 




































ERDC/ITL TR-16-1 viii 


Figure 2.58 Longitudinal pile group responses for the pile group at node 85 and at time 

0.24 sec.52 

Figure 2.59 Rotational pile group response for the pile group at node 85 and at time 0.24 sec.53 

Figure 2.60 Time-history plot of transverse input forces, total force response for all the 

pile groups, and an individual pile group response forces.54 

Figure 3.1 McAlpine alternative flexible approach wall.58 

Figure 3.2 McAlpine flexible approach wall mathematical model.59 

Figure 3.3 (a) Typical 3-D segment of the Impact Deck beam element, (b) Impact force 

applied to the Impact Deck, (c) Typical 3-D beam element.60 

Figure 3.4 Typical 2-D beam element used in lmpact_Deck.60 

Figure 3.5 Plan view of the flexible wall pile layout.61 

Table 3.1 Primary loading curve for the transverse spring model for a McAlpine alternative 
flexible wall bent (3 piles).62 

Figure 3.6 Force time history of Winfield Test # 10.64 

Figure 3.7 Validation of lmpact_Deck against SAP2000 - Transverse displacement at 

node 1.64 

Figure 3.8 Validation of lmpact_Deck against SAP2000 - Transverse displacement at 

node 23.65 

Figure 3.9 Validation of lmpact_Deck against SAP2000 - Transverse displacement at 

node 12’.65 

Figure 3.10 Validation of lmpact_Deck against SAP2000 - Rotation at node 12 and 12’.66 

Figure 3.11 Dynamic transverse response of node 12 and 12’.68 

Figure 3.12 Dynamic response of the rotational spring at node 12 and 12’.68 

Figure 3.13 Dynamic response of the transverse spring located at x = 84.5 ft. .69 

Figure 3.14 lmpact_Deck GUI table of maximum nodal displacements for the McAlpine 

flexible wall.70 

Figure 3.15 Transverse nodal displacement time histories for node 21.70 

Figure 3.16 Longitudinal nodal displacement time histories for node 22.71 

Figure 3.17 Rotational nodal displacement time histories for node 22.71 

Figure 3.18 Transverse wall displacements at 0.252 sec.72 

Figure 3.19 Longitudinal wall displacements at 0.192 sec.72 

Figure 3.20 Rotational wall displacements at 0.22 sec.73 

Figure 3.21 lmpact_Deck GUI table of element minimum and maximum axial forces for 

the McAlpine flexible wall example.74 

Figure 3.22 Axial-force time histories for element 21.75 

Figure 3.23 Axial-force time histories for element 20.75 

Figure 3.24 Shear-force time histories for element 21.76 

Figure 3.25 Shear-force time histories for element 20.76 

Figure 3.26 Moment time histories for element 31.77 

Figure 3.27 Moment time histories for element 21.77 

Figure 3.28 Wall axial forces at 0.2 sec.78 

Figure 3.29 Wall shear forces at 0.220 sec.78 

Figure 3.30 Wall moments at 0.220 sec.79 





































ERDC/ITL TR-16-1 ix 


Figure 3.31 Table of pile group response maximum displacements.79 

Figure 3.32 Response forces for the pile groups at time 0.2200 sec.80 

Figure 3.33 Response forces for the pile groups at time 0.2800 sec.80 

Figure 3.34 Pile group response for the pile group at node 2 and at time 0.298 sec.81 

Figure 3.35 Time-history plot of transverse input forces, total force response for all the 

pile groups, and an individual pile group response forces.82 

Figure 4.1 Guard wall schematic drawing.86 

Figure 4.2 Guard flexible approach wall mathematical model.87 

Figure 4.3 (a) Typical 3-D segment of the Impact Deck beam element, (b) Impact force 

applied to the Impact Deck, (c) Typical 3-D beam element.87 

Figure 4.4 Typical 2-D beam element used in lmpact_Deck.88 

Figure 4.5 Force-displacement relations from push-over analysis of a single guard wall pile.89 

Figure 4.6 Force time history of Winfield Test # 10.91 

Figure 4.7 Validation of lmpact_Deck against SAP2000 - Transverse displacement at 

node 1.91 

Figure 4.8 Validation of lmpact_Deck against SAP2000 - Transverse displacement at 

node 6.92 

Figure 4.9 Dynamic transverse response of node 1.93 

Figure 4.10 Dynamic transverse response of spring at node 6.94 

Figure 4.11 Plastic force-displacement of the transverse spring at node 6.94 

Figure 4.12 lmpact_Deck GUI pile group longitudinal and transverse spring model 

backbone curves.95 

Figure 4.13 lmpact_Deck GUI table of maximum nodal displacements for the guard wall.96 

Figure 4.14 Transverse nodal displacement time histories for node 51.97 

Figure 4.15 Longitudinal nodal displacement time histories for node 51.97 

Figure 4.16 Rotational nodal displacement time histories for node 50.98 

Figure 4.17 Transverse wall displacements at 0.332 sec.98 

Figure 4.18 Longitudinal wall displacements at 0.41 sec.99 

Figure 4.19 Rotational wall displacements at 0.286 sec.99 

Figure 4.20 lmpact_Deck GUI table of element minimum and maximum axial forces for 

the guard wall Example.100 

Figure 4.21 Axial force time histories for element 51.101 

Figure 4.22 Axial force time histories for element 50.101 

Figure 4.23 Shear force time histories for element 51.102 

Figure 4.24 Shear force time histories for element 50.102 

Figure 4.25 Moment time histories for element 51.103 

Figure 4.26 Moment time histories for element 50.103 

Figure 4.27 Wall axial forces at 0.410 sec.104 

Figure 4.28 Wall shear forces at 0.714 sec.104 

Figure 4.29 Wall moments at 0.716 sec.105 

Figure 4.30 Table of pile group response maximum displacements.105 

Figure 4.31 Forces at the three pile group nodes at time 0.332 sec.106 







































ERDC/ITL TR-16-1 


x 


Figure 4.32 Transverse pile group response for the pile group at node 51 and at time 

0.332 sec.106 

Figure 4.33 Time-history plot of transverse input forces, total force response for all the 

pile groups, and an individual pile group response forces.108 

Figure 5.1 Introducing lmpact_Deck.Ill 

Figure 5.2 Geometry for a flexible wall.112 

Figure 5.3 Geometry for a guard wall.113 

Figure 5.4 Geometry for an lmpact_Deck.113 

Figure 5.5 Zooming in the input plot section.116 

Figure 5.6 The zoomed view.116 

Figure 5.7 Selected nodes are highlighted.116 

Figure 5.8 Entering an offset to copy selected nodes.117 

Figure 5.9 Confirming the offset copy (which can be performed multiple times).117 

Figure 5.10 Selected nodes are copied at the offset position.118 

Figure 5.11 Input for an impact time history.119 

Figure 5.12 Extending a time history with 0.0 value.119 

Figure 5.13 Beam properties tab as it appears for a flexible wall.120 

Figure 5.14 Beam properties tab as it appears for an impact deck or guard wall.120 

Figure 5.15 Beam properties tab as it appears for a flexible wall.121 

Figure 5.16 Beam properties tab as it appears for an lmpact_Deck.122 

Figure 5.17 Analyze tab input for analysis method and specified output.124 

Figure 5.18 Selecting finite element nodes where data will be captured.125 

Figure 5.19 Output tab for selecting and viewing select data.126 

Figure 5.20 Output tab with selected data.127 

Figure 5.21 FEO nodal output window showing maximum nodal values for all the nodes.128 

Figure 5.22 Graph of node 70 X-displacement vs time.129 

Figure 5.23 Graph of node 70 Y-displacement vs time.129 

Figure 5.24 Graph of node 70 Z-displacement vs time.130 

Figure 5.25 Animated graph of wall X-displacement.132 

Figure 5.26 Animated graph of wall Y-displacement.132 

Figure 5.27 Animated graph of wall Z-displacement.133 

Figure 5.28 FEO element output window showing maximum and minimum force and 
moments acting on all the elements.134 

Figure 5.29 Graph showing element 60 axial force vs time.135 

Figure 5.30 Graph showing element 60 moment force-length vs time.135 

Figure 5.31 Graph showing element 60 shear force vs time.136 

Figure 5.32 Animated graph of axial forces acting on the wall.137 

Figure 5.33 Animated graph of moments acting on the wall.137 

Figure 5.34 Animated graph of shears acting on the wall.138 

Figure 5.35 Pile Group Response Maximum and Minimum Forces and Moments.139 

Figure 5.36 Pile Group Response Peak Deflections.140 









































ERDC/ITL TR-16-1 xi 


Figure 5.37 Pile Group Response Maximum and Minimum Forces and Moments.140 

Figure 5.38 Animated plot of node 85 X-force and displacement vs time.141 

Figure 5.39 Animated plot of node 85 Y-force and displacement vs time.142 

Figure 5.40 Animated plot of node 85 Z-force and displacement vs time.142 

Figure 5.41 Output run information with selected FEO node output.144 

Figure 5.42 Add node at position 0.0 ft as an inter-monolith node.145 

Figure 5.43 Add interpolated nodes from 3.3541666 ft to 101.4791666 ft.146 

Figure 5.44 Selecting nodes with a left-mouse, click-drag.146 

Figure 5.45 Selected nodes are shown with vertical lines.146 

Figure 5.46 Selected nodes can be copied multiple times with the copy selected nodes 

button.147 

Figure 5.47 The copy selected nodes dialog lets the user specify an offset.147 

Figure 5.48 Select OK the number of times that the selected nodes need to be copied.147 

Figure 5.49 Finally, Add the Final Node.148 

Figure A.l Shape function for axial displacement effect.157 

Figure A.2 Shape function for transverse displacement and rotation effect.159 

Figure A.3 Force-displacement relation of the spring support.164 

Figure B.l Shape function for axial displacement effect.167 

Figure B.2 Shape function for transverse displacement and rotation effect.169 

Figure B.3 Force-displacement relation of the spring support.174 

Figure B.4 Typical McAlpine flexible wall system.175 

Figure B.5 McAlpine flexible wall mathematical model.175 

Figure B.6 Transformation of beam element coordinate system (Local-Global).178 

Figure C.l Shape function for axial displacement effect.181 

Figure C.2 Shape function for transverse displacement and rotation effect.183 

Figure C.3 Force-displacement relation of the spring support.188 

Figure D.l Pipe pile approach wall.189 

Figure D.2 CPGA analytical model.190 

Figure D.3 Simple interaction diagram for 24-inch-diameter pipe pile.191 

Figure D.4 (After Figure 3 Yang 1966) Coefficient of critical buckling strength.193 

Figure D.5 (After Figure 9 Yang 1966) Coefficient decrement of buckling strength.193 

Figure D.6, (After Figure 7 Yang 1966) Coefficient of horizontal load capacity.197 

Figure D.7 (After Figure 2 Yang 1966) Effective embedment of pile at buckling.202 

Figure D.8 Load - displacement plot for pipe pile system.222 

Figure E.l. Plan view of the McAlpine flexible alternative approach wall system.224 

Figure E.2 Relation between Global-Axis and central support Local-Axis.225 

Figure E.3 Location of the Center of Rigidity.225 

Figure E.4 Definition of the forces and distances generated when the pile cap rotate.228 

Figure E.5 Rotational angle definition when the pile cap rotate.229 

Figure G.l Linear variation of acceleration over normal and extended time steps.235 

Figure 1.1 (a) Mass-proportional damping; (b) stiffness-proportional damping.240 











































ERDC/ITL TR-16-1 xii 


Tables 

Table 2.1 Primary loading curve for the transverse spring model of a single pile group with 
a leading vertical pile followed by two batter piles (Lock and Dam 3).38 

Table 2.2 Primary loading curve for the longitudinal spring model of a single pile group 

with a leading vertical pile followed by two batter piles (Lock and Dam 3).38 

Table 2.3 Maximum nodal displacements for the lmpact_Deck example problem.39 

Table 2.4 Extreme forces/moments for the lmpact_Deck example problem.45 

Table 2.5 Transverse forces with respect to time.54 

Table 3.1 Primary loading curve for the transverse spring model for a McAlpine alternative 
flexible wall bent (3 piles).62 

Table 3.2 Primary loading curve for the longitudinal spring model for a McAlpine 

alternative flexible wall bent (3 piles).62 

Table 3.3 Primary loading curve for a spring model for a single 6-ft diameter DIP pile.62 

Table 3.4 Maximum nodal displacements for the McAlpine flexible wall example problem.70 

Table 3.5 Extreme forces/moments for the lmpact_Deck example problem.74 

Table 3.6 Transverse forces with respect to time.82 

Table 4.1 Primary loading curve for the transverse spring model for a bent with two 

vertical piles.89 

Table 4.2 Primary loading curve for the longitudinal spring model for a bent with two 

vertical piles.90 

Table 4.3 Maximum nodal displacements for the guard wall example problem.96 

Table 4.4 Extreme Forces/Moments for the Impact Deck Example Problem.100 

Table 4.5 Transverse forces with respect to time.107 

Table D.l Euler critical buckling load - translating pile top - pinned head condition.195 

Table D.2 Euler critical buckling load - translating pile top - fixed head condition.196 

Table D.3 Euler critical buckling load - translating pile top - pinned head condition.209 

Table D.4 Euler critical buckling load - translating pile top - fixed head condition.210 

Table F.l. HHT-a Method.234 

Table G.l Wilson’s Method: Nonlinear Systems.237 

























ERDC/ITL TR-16-1 


xiii 


Preface 

More than 50% of the U.S. Army Corps of Engineers’ locks and their 
approach walls have continued past their economic lifetimes. As these 
structures wear out, they must be retrofitted, replaced, or upgraded with a 
lock extension. Innovative designs must be considered for the Corps 
hydraulic structures, particularly flexible approach walls, and new tools for 
evaluating these designs must be developed. 

This technical report describes an engineering methodology for the 
dynamic structural response analysis of a flexible approach wall consisting 
of a series of impact beams or impact decks supported by clustered pile 
groups during barge train impact loading. This engineering methodology 
is implemented in a PC-based FORTRAN program and Visual Modeler 
named Impact_Deck, which is also discussed in this report. The 
engineering formulation for Impact_Deck uses an impact-force time 
history acting on the clustered pile group founded flexible impact beams 
or impact decks to characterize the impact event. This impact-force time 
history may be obtained using a companion program Impact_Force 
(Ebeling et al. 2010). 

This report was authorized by Headquarters, U.S. Army Corps of 
Engineers (HQUSACE), and was written from October 2013 to March 
2014. It was published under the Navigation Systems Research Program, 
Work Unit “Flexible Approach Walls.” Jeff McKee was the HQUSACE 
Navigation Business Line Manager. 

The Program Manager for the Navigation Systems Research Program was 
Charles Wiggins, Coastal and Hydraulics Laboratory (CHL), U.S. Army 
Engineer Research and Development Center (ERDC). Jeff Lillycrop was 
Technical Director, CHL-ERDC. The research was led by Dr. Robert M. 
Ebeling, Information Technology Laboratory (ITL), ERDC, under the 
general supervision of Dr. Reed L. Mosher, Director, ITL-ERDC; Patti S. 
Duett, Deputy Director, ITL-ERDC. This work effort was also done under 
the general supervision of Dr. Robert M. Wallace, Chief, Computational 
Science and Engineering Division (CSED), ITL, during the initial 
formulation and programming stage. During the data interpretation and 
report writing stages, Elias Arredondo, Dr. Kevin Abraham, and 
Dr. Jerrell R. Ballard were Acting Division Chiefs. Dr. Ballard is the CSED 



ERDC/ITL TR-16-1 


xiv 


Chief for the final stage of the publication process. Dr. Ebeling was the 
Principal Investigator of the “Flexible Approach Walls” work unit. 

This report was written by Barry C. White of ITL-ERDC, Professor Jose 
Ramon Arroyo, University of Puerto Rico at Mayaguez, and Dr. Ebeling of 
ITL-ERDC. White is with the Computational Analysis Branch (CAB), of 
which Elias Arredondo is Chief. 

Impact_Deck software formulation was developed by Arroyo and Ebeling. 
Input specifications for the pertinent engineering features and boundary 
conditions of the three different wall types to be analyzed by this software 
were provided by Ebeling, White, and Arroyo. Programming for the 
engineering formulation was led by Arroyo, with support from White and 
Ebeling. The Graphical User Interface (GUI) comprised of the Visual 
Modeler and an extensive and detailed Output Visualization software 
package was developed by White with support from Arroyo and Ebeling. 
Example problem input definition was provided by Ebeling, Arroyo, and 
White. Initial engineering program validation and example problem 
engineering assessments were led by Arroyo. Initial engineering 
formulation documentation was created by Arroyo, with support by 
Ebeling and White. Example problem output was collected and interpreted 
by White and Arroyo. White was lead on the final organization of this 
report, including the compilation, recasting, and reduction of the 
engineering formulation description and example problems using the 
Impact_Deck processor, with the aid of Arroyo. Having provided the 
Visual Pre- and Post-Processor for Impact_Deck, White led in writing the 
user interface and output visualization sections for the report. White also 
provided additional example problems for verification and to support 
observations highlighting the unique engineering advantages of each of 
the three flexible approach wall systems, with the aid of the Visual Pre- 
and Post-Processor. These unique features were interpreted from the data 
results by Ebeling and White. 

At the time of publication, COL Bryan S. Green was Commander, ERDC, 
and Dr. Jeffery P. Holland was the Director. 



ERDC/ITL TR-16-1 


xv 


Unit Conversion Factors 


Multiply 

By 

To Obtain 

cubic feet 

0.02831685 

cubic meters 

cubic inches 

1.6387064 E-05 

cubic meters 

cubic yards 

0.7645549 

cubic meters 

degrees (angle) 

0.01745329 

radians 

feet 

0.3048 

meters 

foot-pounds force 

1.355818 

joules 

inches 

0.0254 

meters 

inch-pounds (force) 

0.1129848 

newton meters 

knots 

0.5144444 

meters per second 

microns 

1.0 E-06 

meters 

miles (nautical) 

1,852 

meters 

miles (U.S. statute) 

1,609.347 

meters 

miles per hour 

0.44704 

meters per second 

pounds (force) 

4.448222 

newtons 

pounds (force) per foot 

14.59390 

newtons per meter 

pounds (force) per inch 

175.1268 

newtons per meter 

pounds (force) per square foot 

47.88026 

pascals 

pounds (force) per square inch 

6.894757 

kilopascals 

pounds (mass) 

0.45359237 

kilograms 

pounds (mass) per cubic foot 

16.01846 

kilograms per cubic meter 

pounds (mass) per cubic inch 

2.757990 E+04 

kilograms per cubic meter 

pounds (mass) per square foot 

4.882428 

kilograms per square meter 

pounds (mass) per square yard 

0.542492 

kilograms per square meter 

slugs 

14.59390 

kilograms 

square feet 

0.09290304 

square meters 

square inches 

6.4516 E-04 

square meters 

tons (force) 

8,896.443 

newtons 

tons (force) per square foot 

95.76052 

kilopascals 

tons (2,000 pounds, mass) 

907.1847 

kilograms 

tons (2,000 pounds, mass) per square foot 

9,764.856 

kilograms per square meter 

yards 

0.9144 

meters 




ERDC/ITL TR-16-1 


1 


1 Dynamic Structural Time-History 

Response Analysis of a Flexible Approach 
Wall Supported by Clustered Pile Groups 
during Impact with a Barge Train 

1.1 Introduction - Glancing Impact Blows and Flexible Approach 
Wall Structural Systems 

A glancing blow impact event of a barge train impacting an approach wall as 
it aligns itself with a lock is an event of short duration. The contact time 
between the impact corner of the barge train and the approach wall can 
range from one second to several seconds. In order to reduce construction 
costs as well as to reduce damage to barges during glancing blow impacts 
with lock approach walls, the next generation of Corps approach walls are 
more flexible than the massive, stiff-to-rigid structures constructed in the 
past. A flexible approach wall or flexible approach wall system is one in 
which the wall/system has the capacity to absorb impact energy by 
deflecting or “flexing” during impact, thereby affecting the dynamic impact 
forces developing during the impact event. Pile-founded approach wall 
structural systems are characterized as flexible structures. This report 
summarizes an engineering methodology as well as the corresponding 
software for performing a dynamic structural response analysis. The 
analysis is of a flexible impact approach wall deck or impact beam system 
supported by piles representing using an elastic-plastic spring model of 
each of the clustered pile groups to a barge train impact event. The PC- 
based software used for conducting the dynamic structural response 
analysis is referred to as Impact_Deck. A pulse-force time history of the 
barge train impacting the flexible approach wall is used to dynamically load 
the model. The PC-based software Impact_Force (Ebeling et al. 2010) is 
used to develop the pulse-force time history required by the Impact_Deck 
software. 

1.2 Examples of the Next Generation Flexible Approach Walls 

1.2.1 Lock and Dam 3 

Lock and Dam No. 3 (Lock and Dam 3) is a lock and dam located near Red 
Wing, Minnesota on the Upper Mississippi River around river mile 796.9. 




ERDC/ITL TR-16-1 


2 


It was constructed and placed in operation July 1938. The site underwent 
major rehabilitation from 1988 through 1991. In recent years, a guide wall 
extension was added to the project. 


The structure consists of eight reinforced concrete monoliths of 104 ft 10 
in. each. It was constructed by joining eight concrete blocks of 12 ft 6 in. 
with a free end of 37.5 in. at both ends. Each block is supported over two 
rows of piles with three piles per row. That results in a monolith supported 
by 48 piles, where 36 of these piles are battered with a batter of 1:4. Over 
the piles, the monolith’s dimensions are 5 ft in height and 22 ft in width. 
One end of the deck is pinned and connected to a massive circular concrete 
cell and the other end is free. The inter-monolith connections are 
connected by using four reinforcing steel bars just to transfer the axial and 
shear force and no flexural moment transfer. The piles have a diameter of 
2 ft, and are symmetrically located at each concrete block. The weight of 
each precast concrete beam is approximately 6,919 tons. Some of the 
drawings/plans of this flexible impact deck are shown in Appendix D. 
Examples of this type of Corps flexible approach wall is shown in 
Figures 1.1 and 1.2 for Lock and Dam 3. 


Figure 1.1 Front-end cells of Lock and Dam 3. 





























































ERDC/ITL TR-16-1 


3 


Figure 1.2 Lock and Dam 3 cross-section and plan view. 



1.2.2 McAlpine flexible wall 

McAlpine Locks and Dam are located in downtown Louisville, Kentucky. 
The dam is at mile 604.4 of the Ohio River and the locks are in the 
Louisville and Portland Canal on the Kentucky side of the river. The 56 ft x 
600 ft auxiliary lock was completed in 1921. The no ft x 1200 ft main 
chamber opened in 1961. A new lock chamber (110 ft x 1200 ft) began 
operation in 2009. 

The alternative flexible approach wall structure discussed in this section 
consists of a continuous elastic concrete beam with segments spanning 
approximately 96 ft in length. The continuity of the beam is achieved by 
means of shear key at each pile group support. That means the beams just 
transfer the longitudinal and transverse forces with no moment transfer at 
each pile support. The axial and transverse forces at the end of the beams 
are transferred to the pile cap by means of a shear key. The shear key has a 
length of 11.5 ft. The length of the shear key is the distance between two 
consecutive concrete beams. The shear key is part of the massive pile cap 
that rests over the pile group. The pile group consists of three piles, each 
one with a diameter of 5 ft 8 in. They are arranged in a triangular scheme 
to absorb the torsion generated at the pile group due to the eccentricity 
between the center of the pile group and the location of the end of the 
beams that rest over the pile group. A plan view drawing and a cross- 
section view are presented in Figure 1.3. 





































ERDC/ITL TR-16-1 


4 


Figure 1.3 McAlpine flexible walls. 



1.2.3 Guard walls 

This kind of flexible approach wall can be found at numerous Corps locks. 
The structure consists of a continuous elastic concrete beam with a span of 
approximately 50 or 60 ft long, each segment. The continuity of the beam 
is achieved by means of shear key at each pile-group support. That means 
the beams just transfer the longitudinal and transverse forces with no 
moment transfer at each pile support. The axial and transverse forces at 
the ends of the beams are transferred to the pile cap by means of a shear 
key. The shear key is a concrete block between the end and start of two 
consecutive flexible beams constraining the motion of the beam to the 
motion of the pile bent. The length of the shear key is equal to the width of 
the pile cap of the pile group. The shear key is part of the massive pile cap 
that rests over the pile group. The pile group consists of two aligned piles, 
each with a diameter of 5 ft 8 in. The two piles are arranged in such a way 
that no torsion is transferred to the pile group. A plan view drawing and a 
cross-section view are presented in Figure 1.4. 

1.3 Overview of Dynamic Time-History Response Analysis of a 

Flexible Beam Supported Over Elastic-Plastic Spring Supports 

Due to the flexibility and the mass of the new generation flexible approach 
walls, the impact event can be a dynamic event from the point of view of the 
mathematical structural model. In structural dynamics the mathematical 













































ERDC/ITL TR-16-1 


5 


model of bodies of finite dimensions undergoing translatory motion are 
governed by Newton’s Second Law of Motion, expressed as 

^F = m»a (1.1) 

where F are forces, m is mass, and a is acceleration at each time step t 
during motion. 


Figure 1.4 Guard wall schematic drawing. 



In the mathematical model, the forces acting on the flexible wall mass at 
each time step t are (l) the impact force at time step t, (2) the elastic 
restoring forces (of the beam), and (3) the damping forces (of the beam). 
This report discusses an engineering methodology that uses Equation 1.1 
to compute the dynamic structural response of a flexible impact beam 
supported over flexible supports of the mathematical model to the impact- 
force time history. The impact event is idealized as shown in Figure 1.5 for 
the mathematical beam and impact event model of the force time history, 
Fnormal-wa u(t) is developed by scaling of existing pulse-force time histories 
recorded during the full-scale barge impact experiments conducted at 
Winfield Lock & Dam (Ebeling et al. 2010) and the Pittsburgh Prototype 
tests (Patev et al. 2003). The force time history shown in Figure 1.6 
denotes the component of the pulse force time history acting normal to the 
wall. Initial barge train contact with the wall starts at time ti and ends at 
time £2. The solution to this dynamic problem will be a succession of 
solutions at user-specified time steps starting at time ti. These solution 
time steps are dictated by the time step the user selects for the barge train 
impact-force time history created using the PC-based software 
























































ERDC/ITL TR-16-1 


6 


Impact_Force (Ebeling et al. 2010). Due to the nature of dynamic 
structural response of some types of beams with consideration of both the 
duration and frequency characteristics of the impact-force time history, 
the peak response of the simply supported, flexible impact beam may 
occur during impact (i.e., between times ti and t 2 ) or after impact 
concluded (i.e., after time t 2 ). The engineering methodology discussed in 
this report and corresponding software are capable of addressing both 
types of dynamic structural systems responses. 

Figure 1.5 Barge train impacting at a fixed impact position 
along the simply supported, flexible impact beam mathematical 
model with the barge train oriented at an approach angle 0 to 
the wall’s Xe/obai axis (plan view). 



Figure 1.6 Example of an impact pulse-force time history. 


























ERDC/ITL TR-16-1 


7 


An alternative formulation incorporated within the PC-based program 
Impact_Deck allows for the specification of a barge train having an initial 
contact with the impact beam at a position designated X_Impact (in PC- 
program Impact_Deck input terminology) at time ti and moving in contact 
at a constant velocity (V) along the beam as shown in Figure 1.7. The 
position of the impact point force moves with time after contact. The time 
after contact is designated as an increment in time Ati, and it occurs at an 
absolute time of [ti + Ati]. The change in position of the contact force is 
designated a distance AX from initial contact point X_Impact and is given 
by 


AX'= 7 • At,. 


( 1 . 2 ) 


Figure 1.7 Barge impact point force moving along the wall from initial contact time ti to final contact time fa. 


Flexible 



The position of the point load along the beam at time ti is 

[Xpoint load] t; =^- Im P aCt +A* (1.3) 


Substituting from Equation 1.2, Equation 1.3 becomes 














ERDC/ITL TR-16-1 


8 


[Xpoint loadlt, =[X_Impact +V.At ( ] (1.4) 

Thus, the normal force time history of Figure 6 moves along the beam 
from time ti to time t 2 according to the user-specified velocity ( V). 

1.4 Report Contents 

The engineering methodology discussed in this report is implemented in a 
PC-based FORTRAN program named Impact_Deck, which is also discussed 
in this report. A pulse-force time history (normal to the flexible approach 
wall) is used in this dynamic time-history response analysis to represent the 
demand made of the flexible beam supported over elastic-plastic springs 
during the impact event. The impact-force time history to be used in the 
Impact_Deck analysis is created by the companion PC-based program 
Impact_Force (Ebeling et al. 2010). The engineering formulation for 
Impact_Force uses the impulse momentum principle to convert the linear 
momentum of a barge train into a pulse-force time history acting normal to 
the approach wall. 

The engineering formulation developed for and implemented in 
Impact_Deck assumes that the District engineer will have knowledge of 

1. Length, modulus of elasticity, cross-sectional area, moment of inertia and 
mass per unit length (equal to the weight per unit length divided by the 
gravitational constant, including hydrodynamic added mass for the beam) 
of the flexible impact beam 

2. Point of initial impact 

3. Velocity (V) the barge train moves along the approach wall during impact 

4. Dynamic coefficient of friction between the wall and the barge train 

5. Force-displacement relationship of a pile group 

6. Impact pulse-force time history normal to the impact beam 

The engineering formulation developed for Impact_Force assumes that 
the District engineer will have knowledge of 

1. Size, the weight (and mass) of the barge train (including hydrodynamic 
added mass) 

2. Barge approach velocity (often expressed in local barge coordinates) 

3. Approach angle (the angle measured from the face of the wall to the side of 
the barge train) 



ERDC/ITL TR-16-1 


9 


This information will be required for the usual, unusual, and extreme 
design load cases. 

Sections 2-4 describe the relationships that comprise the engineering 
formulation used to solve Newton’s Second Law of Motion for the dynamic 
response of a flexible impact beam supported on groups of piles and 
subjected to a barge train impact. The groups of piles are modeled as stiff 
elastic-plastic springs. The barge train impact is modeled using a 
representative pulse-force time history applied normal to the point of 
contact between the barge train and the flexible approach wall. In this 
initial engineering formulation implemented in Impact_Deck, the 
numerical solution of Newton’s Second Law of Motion for the impact 
pulse-force time history is applied to the flexible impact beam. The 
numerical solution makes use of the Wilson’s 0 method to solve the 
equations of motion of the multiple degrees of freedom (MDOF) system in 
the time domain. Each section discusses the results of this analysis for a 
different structural system. The analysis in section 2 is for the Lock and 
Dam 3 guide wall structural system; section 3 is for the McAlpine flexible 
wall structural system; and section 4 is for a typical guard wall structural 
system. 

Section 5 introduces the user to the Graphical User Interface (GUI), 
Engineering Processor, and Visual Post-Processor named Impact_Deck. 
An example problem presents the features for input and the output 
visualization. 

Section 6 presents the conclusions of this report based on the 
Impact_Deck computer software. 

Appendix A discusses the formulation for a structural impact deck founded 
on rows of pile groups. This type of flexible approach wall structural system 
was used for the Lock and Dam 3 approach wall extension. Each pile group 
cluster consists of an in-line row of vertical and batter piles. 

Appendix B discusses the formulation for a flexible approach wall founded 
on a triangular formation of three vertical piles modeled using 
Impact_Deck. The McAlpine lock alternative flexible approach wall 
structural system is an example of this type of flexible-impact structure. 

Appendix C discusses the formulation for a simply supported flexible- 
impact beam supported by two groups of clustered vertical piles modeled 



ERDC/ITL TR-16-1 


10 


using Impact_Deck. Guard walls are an example of this type of flexible- 
impact structure. Each pile-group cluster consists of an in-line row of 
vertical piles. 

Appendix D discusses the formulation for a two-translational spring model 
of a row of piles that is used in Impact_Deck. This appendix discusses the 
non-linear, force-deflection relationship for primary loading and for 
unload-reload response of the clustered group of piles using the push-over 
method of analysis and CPGA 1 software applied to the Lock and Dam 3 
approach wall extension problem. 

Appendix E discusses the formulation for a single rotational spring model 
of a clustered group of vertical piles. This type of flexible approach wall 
structural system, proposed for use at McAlpine Locks and Dam, is 
referred to as the “alternative” flexible approach wall system. 

Appendix F discusses the numerical method formulation for the time 
integration of the Equation of Motion by HHT-a. 

Appendix G discusses the numerical method formulation for the time 
integration of the Equation of Motion by the Wilson -0 method. 

Appendix H summarizes the member end release details used in 
Impact_Deck for a load applied at the end release node. 

Appendix I summarizes the Raleigh Damping formulation for the three 
flexible approach walls. 

Appendix J summarizes the Impact_Deck ASCII input file. 

Appendix K lists key Impact_Deck FORTRAN program variables. 


1 CPGA is CASE software for the Pile Group Analysis. 



ERDC/ITL TR-16-1 


11 


2 Impact Deck Approach Wall - Lock and 
Dam 3 Example 

2.1 Introduction 

This chapter summarizes an engineering methodology using the 
Impact_Deck software for performing a dynamic structural response 
analysis of a flexible impact deck supported over groups of clustered piles 
and subjected to a barge train impact event. This example is based on the 
real-world example at Lock and Dam 3 on the Mississippi River in 
Minnesota. 

This example models a glancing blow impact event of a barge train 
impacting an approach wall as it aligns itself with a lock is an event of short 
duration; the contact time between the impact corner of the barge train and 
the approach wall can be as short as a second or as long as several seconds. 
The next generation of Corps approach walls is more flexible than the 
massive, stiff-to-rigid structures constructed in the past in order to reduce 
construction costs as well as to reduce damage to barges during glancing 
blow impacts with lock approach walls. A flexible approach wall or flexible 
approach wall system is one in which the wall has the capacity to absorb 
impact energy by deflecting or “flexing” during impact, thereby affecting the 
dynamic impact forces that develop during the impact event. 

2.2 Lock and Dam 3 - Physical Model 

Lock and Dam 3 is located in Welch Township, Minnesota, on the Upper 
Mississippi River, approximately 6 miles upstream from Red Wing, 
Minnesota (around river mile 796). 

Artist renderings, idealized cut-away sections, and pictures of Lock and 
Dam 3 are shown in Figures 2.1 through 2.12, as well as in Figures D.i and 
D.2 (Appendix D). The impact deck of the flexible guide wall consists of 
eight reinforced concrete monoliths. Each monolith is approximately 
104 ft 10 in. in length 1 . Each impact-deck monolith was constructed by 
joining together eight concrete precast block segments that are 12 ft 6 in. 
in length. Each 12 ft 6 in. block segment is supported over two rows of 


1 The cited dimensions and those shown on construction drawings may deviate slightly from as-built 
conditions. 



ERDC/ITL TR-16-1 


12 


piles, with three piles per row. The front row of piles is vertical and the 
back two rows of piles have a batter of 1:4. The piles are concrete-filled 
pipe piles having a diameter of 2 ft and a vertical height of approximately 
72 ft (Figure D.i). Figure 2.6 shows one of the pre-cast 12 ft 6 in. bases for 
each block segment prior to installation. The base is hung from the top of 
the piles (Figure 2.8). The impact-deck monolith is then cast on top of 
eight neighboring block segments. Figure 2.2 shows the relationship of the 
two and a half precast bases (the areas above the piles but below the 
monolith, separated by vertical lines) to the overtopping monolith and 
supporting pile groups. That results in an impact-deck monolith being 
supported by 48 flexible piles, where 16 of these piles are vertical and 
another 32 of these piles are battered. Over the piles, the impact-deck 
monolith has dimensions of 5 ft height and 22 ft width. 

Each end of adjoining monoliths is structurally detailed to provide for 
shear and axial load transfer but no moment transfer between monoliths 
using shear bars that are close to the line through the cross-section center. 
The ends of each monolith are 37.5 in. from the last supporting pile group. 

One end of the flexible guide wall impact deck (consisting of eight 
monoliths) is pin-connected to a massive circular concrete cell and the other 
end, adjacent to the existing approach wall, is free. The weight of impact 
deck (all eight monoliths) is approximately 627,655 kg (6,919 tons). 


Figure 2.1 Front end cell of Lock and Dam 3. 






















































ERDC/ITL TR-16-1 


13 


Figure 2.2 Pinned connection between the circular cell and the Impact Deck at Lock and Dam 3. 



Figure 2.3 Arrangement of piles at Lock and Dam 3. 













































































































ERDC/ITL TR-16-1 


14 


Figure 2.4 Impact Deck supported over piles at Lock and Dam 3. 



Figure 2.5 An upstream view starting at the concrete 
cell of the pile-founded flexible approach wall at Lock 
and Dam 3. 




















ERDC/ITL TR-16-1 


15 


Figure 2.6 Precast bases before installation. 



Figure 2.7 Precast bases during installation. 
















ERDC/ITL TR-16-1 


16 


Figure 2.8 Precast bases connected to piles. 



Figure 2.9 Construction joint between concrete block segments (axial, shear 
and moment transfer connection). 





















ERDC/ITL TR-16-1 


17 


Figure 2.10 Construction joints. 



Figure 2.11 Massive concrete circular cell. 




















ERDC/ITL TR-16-1 


18 


Figure 2.12 Pile-founded flexible impact deck 
structure. 



2.3 Lock and Dam 3 - Construction Drawings 

In Figures 2.13-2.18, some general construction drawings are provided 
that demonstrate the pile arrangement, the inter-monolith arrangement, 
and circular concrete cell location. Figure 2.14 shows the plan view and 
pile layout for the concrete cell at the start of the guide wall. Figure 2.5 is a 
picture of this completed concrete cell. A typical cross section of the 
flexible impact deck structural system is presented in Figure 2.15. This 
figure also shows a close-up, plan view of a 12 ft 6 in. long precast base 
segment. Figure 2.6 is a picture of this precast base segment. 

2.4 Lock and Dam 3 - Mathematical Model 

Lock and Dam 3 can be considered as a beam element because its length is 
much greater than the other two directions. The length is 838.66 ft and the 
height and width are 5 ft and 22 ft, respectively. The model can be seen in 
Figure 2.19. 







ERDC/ITL TR-16-1 


19 


Figure 2.13 Lock and Dam 3 guide wall plan view 1 of 2. 



iii 


SNVId 


eiosauujw 'ipRM 
lieMapjns wea >8 >pcr| 










































































ERDC/ITL TR-16-1 


20 


Figure 2.14 Lock and Dam 3 guide wall plan view 2 of 2. 


















































































ERDC/ITL TR-16-1 


21 


Figure 2.15 Lock and Dam 3 guide wall detail of pile layout and end cell. 




























































































ERDC/ITL TR-16-1 


22 


Figure 2.16 Lock and Dam 3 guide wall section view. 






































































































































































































































ERDC/ITL TR-16-1 


23 


Figure 2.17 Lock and Dam 3 guide wall plan - 5. 







































































































































































































































ERDC/ITL TR-16-1 


24 


Figure 2.18 Lock and Dam 3 guide wall plan - 6. 






























































































































ERDC/ITL TR-16-1 


25 


The mathematical model can be done using 3-D beam elements. A 3-D 
beam element has 6 degrees of freedom per node, producing 12 degrees of 
freedom per element. The degrees of freedom per node are 3 translations 
and 3 rotations as seen in Figure 2.20. The applied normal force F x (t) is 
the impact-force time history developed using the PC-based software 
Impact_Force. The applied parallel force F y (t) is a (decimal) fraction of the 
normal force calculated using the dynamic coefficient of friction between 
the barge and the impact deck surfaces. Note the armor rubbing surfaces 
cast into the impact deck face can be seen in detail in Figure 2.10. 


Plan View 


Figure 2.19 Lock and Dam 3 flexible approach wall. 

Lock and Dam 3 Flexible Approach Wall Extension 

838'-6 1/8" 


Pile 

Founded 
Concrete 
Filled Cell 


Barge Train Impact 

V 


Pile Group 
Supported 
Monoliths 



Existing 

Approach 

Wall 


1.92" Expansion Joint 
ShearConnection 
(no moment transfer) 

2" Expansion Joint 
ShearConnection 
(no moment transfer) 


1.92" Expansion Joint 
(no shear or moment transfer) 


Figure 2.20 (a) Typical 3-D segment of the impact-deck beam element, (b) Impact force 
applied to the Impact Deck, (c) Typical 3-D beam element. 







































































ERDC/ITL TR-16-1 


26 


If the model used to describe the beam is developed in the plane, the beam 
element has 3 degrees of freedom per node and 6 degrees of freedom per 
element. The degrees of freedom per node are 2 translations and 1 
rotation, as seen in Figure 2.21. Based on the notation of Figure 2.20, the 
force and moment conditions for node i are Fi, x = Vi, Mi, x = o, Fi, y = Fi, Mi, y 
= o, Fi, z = o, and Mi, z = Mi, and for node/are Ff x = Vf, Mf, x = o, Ff, y = Ff, Mf y 
= o, Ff, z = o, and Mf z = Mf. Basically, to transform a 3-D beam element to a 
2-D (plane element), the moment about the “x” axis, the moment about 
the “y” axis, and the force in the “z” directions are equal to zero. 


Figure 2.21 Typical 2-D beam element used in lmpact_Deck 



The Impact_Deck PC-based computer program is based on beam elements 
loaded and deformed in the plane “y-x”. The mathematical model is 
presented in Figure 2.22. The connection of the impact deck to the 
concrete circular cell is assumed as pinned and the end of the last 
monolith is assumed to be free. The inter-monolith connection is 
considered as an internal pin where the moment is zero (i.e., no moment 
transfer). The normal and shear impact force (time history) is applied 
along the beam elements and has a variation in time and position. The 
load moves with a constant velocity, so the position of the load varies 
linearly with respect to time. The impact deck is supported by equally 
spaced nonlinear springs in the global “x” and “y” directions. These 
nonlinear springs represent the reaction provided by each row of three 
piles through soil-structure interaction with the foundation soil(s). 

A description of the beam element used for the first element (i.e., element 
that is pinned and connected to the rigid circular cell) is presented in 
Figure 2.23. At node 1, which is at the cell to monolith number 8 interface 
(left side in Figure 2.19), the element has zero displacements (for these two 
DOFs) and also, a zero bending moment. The other force, moment, 
displacement, and rotations are not equal to zero. 











ERDC/ITL TR-16-1 


27 


Figure 2.22 Lock and Dam 3 mathematical model. 


Support at circular 
rigid structure 

L|.' 

£ 

L 

y, v(y,t), t F x 

Inter-monolith pin 
connection 

i 

Fy 

j™ j— 

b[a[a[a[a[a[a[a[a[a[a[a[a[a[a[a[b 

T 

1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 ’ 

Typical first monolith 

f 

Inter-monolith pin ^_ 

connection 


_^ Inter-monolith pin 

connection 

i 

i 

^ r r r r r r r r r r r r r r r 

b[ a [a [a [a [a [a [a [a [a [a [a [a [a [a [a [b 


1111111111111111' 

Typical internal monolith 

1 

Inter-monohth pin ^_ 

connection 

• 

j 


_^ Beam with 

free end 

|T r r r r r r r r r r r r r r f 

Pi a [a [a [a [a [a [a [a [a [a [a [a [a [a [a [l 

’] 

1 

1111111111111111 

Typical last monolith 

1 


Figure 2.23 First beam element used in Lock and Dam 3 mathematical model. 


A 

_ 

\w, = 0, V, 

P 6 h M, = 0 

0 2 , M: Y 

\ w 

Y 

? , V 2 

^ 

Node 1 = Upstream 

^ w 

Hi = 0 Fi ± _ 

— W 

L 

»i,F, 

Node 2= Downstream 

^ b 

Nodel 

Node 2 



A description of the beam element used for the last element (i.e., element 
adjacent to the existing approach wall) is presented in Figure 2.24. The 
force, moment, rotation, and displacement at node (n-i) are similar to node 
i of the typical beam element. At node n, located at the monolith number 1 
to the existing approach wall interface on the right side in Figure 2.19, the 
forces and moment (at node n) are equal to zero (i.e., free beam end). 


Figure 2.24 Last beam element used in Lock and Dam 3 mathematical model. 


Ui> F l ■ 


A w,-, Vi 

A\ 

-> 


0 i,M t 


+- 


A WfVf -0 

6 f , M f = 0 \ u f F f =0 

-> 

■+ 


Node (n-1) 


Noden 


Node (n-1) = Upstream 
Noden = Downstream 




















































ERDC/ITL TR-16-1 


28 


A description of the beam element used for the inter-monolith pin 
connection is presented in Figure 2.25. At the inter-monolith pin 
connection (node s), the bending moment is equal to zero; but the other 
forces, moments, displacements, and rotations are not equal to zero. 


Figure 2.25 Description of two beam elements connected at the inter-monolith connection. 


A K 

0 r , M r 


Q s , M s = 


1 


-> 


_ _ w s , V s 
= ° u s ,F s 

= 4 —> 


Noder 


Nodes 


Noder = Upstream 

Nodes = inter-monolith pin 
connection 

Node t = Downstream 


A w s ,V s 


u s ,F s 


G S ,M S =0 


A w t> v t 

G t ,M t 'f\ Uu p t 

i —> 
-+■ 


Nodet 


Figure 2.26 shows a typical internal monolith and the locations of the 
connected inter-monolith beam elements. Observe that the individual pile 
group row number (i.e., at the location of each pair of springs) is also 
shown in this Figure and labeled as numbers 1 through 16. 


Figure 2.26 Beam elements numbering in a typical internal monolith. 



The behavior of a single row of three piles under static lateral load was 
conducted to determine the force-versus-displacement relationship for the 
nonlinear springs used in the Lock and Dam 3 model. A description of the 
cross-section of Lock and Dam 3 is presented in Figure 2.27 with the three 
batter piles that resist the lateral load. The system has two battered piles 
with an inclination of 1 horizontal to 4 vertical and one vertical pile. The 
piles extend 24 ft above ground surface and are embedded a vertical 
distance of 48 ft into the ground. Each pipe pile is 24 in. in diameter and is 
filled with concrete. 




























































ERDC/ITL TR-16-1 


29 


Figure 2.27 Cross-section of Lock and Dam 3. 



Two separate push-over analyses were conducted to define the pair of 
nonlinear impact deck springs representing the soil-structure interaction 
responses in the transverse and longitudinal directions of an individual 
batter pile group using the computer program CPGA. The transverse push¬ 
over result shown in Figure 2.28a is consistent with those reported in the 
Ebeling et al. (2012) Figure 4.4 (or Figure B.8). The resulting nonlinear 
force-versus-displacement relationship for loading in the transverse 
direction is a result of the development of plastic hinging at various 
locations within the individual piles, and of the development of (soil-to- 
pile interface) tension and/or pile tip (end) bearing soil failures of 
individual piles within the batter pile system as the applied lateral load is 
increased on the single batter pile group. Details of this push-over analysis 
are discussed in section 4 of Ebeling et al. (2012). Figure 2.28a, force- 
versus-displacement curve (with a white background, shows that a 
horizontal force of 180 kips results in a lateral displacement of 0.2 ft 
(2.46 in.); a force of 250 kips results in a lateral displacement of 0.41 ft 
(4.9 in.); a force of 272 kips results in a lateral displacement of 0.63 ft 
(7.5 in.); and a force of 317.75 kips results in a lateral displacement of 1.1 ft 
(13 in.). The data describing this curve is shown as a table in this user 
interface (Figure 2.28a). The curve used to define the elastic loading phase 
of the load-unload nonlinear process that will be discussed in section 2.5. 



























































ERDC/ITL TR-16-1 


30 


Figure 2.28 Transverse and longitudinal push-over results for a 
single row of three piles aligned in the transverse direction. 



a) Transverse direction 



b) Longitudinal direction 

The longitudinal push-over results are shown in Figure 2.28b. The 
resulting nonlinear force-versus-displacement relationship is a result of 
the development of plastic hinging, first occurring at the pile cap and then 
at a point along the pile located below the mud line within the individual 
piles of the three-pile group as the lateral load is applied to the single 
batter pile group in the longitudinal direction (i.e., perpendicular to the 
line of piles). The Figure 2.28b force-versus-displacement curve (with a 
white background) shows a horizontal force of 123 kips resulting in a 
lateral displacement of 0.35 ft (4.18 in.), and a force of 149.7 kips resulting 
in a lateral displacement of 0.64 ft (7.6 in.). The curve used to define the 




































































ERDC/ITL TR-16-1 


31 


elastic loading phase of the load-unload nonlinear process will be 
discussed in section 2.5. 

The ultimate push-over capacity of the pile group in the transverse 
direction is 317.75 kips as compared to a capacity of 149.7 kips in the 
longitudinal direction, a factor of 2.1. For the same level of horizontal 
displacement, the push-over curve in the longitudinal direction is of lower 
magnitude force than in the transverse direction. This is because for 
longitudinal loading each of the three piles responds more like “vertical” 
piling under a lateral load. Vertical pile push-over response behavior is 
discussed in detail in section 3 of Ebeling et al. (2012). It is observed that 
vertical pile behavior under lateral loading occurs without the soil-to-pile 
failures and the “pole-vaulting” actions that are unique to a batter pile 
group subjected to an applied line of loading in the transverse direction 
(i.e., in line with the three-pile group). These push-over results are further 
discussed in section 2.9. 

2.5 Nonlinear force-deflection relationship for the spring supports 

Impact_Deck has the capability to calculate the response of the spring 
supports during the time-history analysis even if the springs possess plastic 
behavior in their force-displacement relationship of an individual pile 
group. The spring is considered as “linear” if the load in the spring is below 
the elastic displacement Seias and the elastic force Feias as shown in 
Figure 2.29. If the load is reduced, and the force-displacement is below 
point 1, the unloading path follows the same path as the previous loading 
phase. The loading phase in this figure is depicted as the green arrows and 
the unloading phase by the red arrows. However, if the load is greater than 
the elastic displacement and is in the loading stage, the load will follow the 
path shown using the green arrows until it reaches the maximum force- 
displacement value, labeled as point 2 in this figure. If unloading occurs 
from this point, it will unload following the user-specified slope that follows 
the unload path starting at point 2 and moves in the direction of point 4. If 
the pile group nonlinear “spring” is never again subjected to a force as large 
as that corresponding to point 2 on this figure, the force-displacement 
response will remain along a line from points 2 to 4, until zero force is 
reached. However, a plastic permanent deflection equal to the distance from 
the origin to point 4, lateral displacements will result for the pile group. 
Another scenario is if the load should increase again and go above the point 
2 force magnitude, the load-displacement response will follow along the 
“original backbone curve” moving from point 2 towards point 3. If the force 



ERDC/ITL TR-16-1 


32 


reaches a maximum value somewhere between point 2 and 3 and starts to 
decrease again, the load-deflection will follow the same unload slope as the 
slope between points 2 and 4, but start from the new maximum force- 
deflection point. Lastly, if the force-deflection value is greater than that 
corresponding to point 3, Impact_Deck assigns a zero value to this spring 
because the maximum force value was reached and failure will occur. 


Figure 2.29 Force-displacement relation of the spring 
support. 



2.6 Solving for the motion of the structure 

The equations of motion for a flexible approach wall structure comprised of 
decks supported on clustered pile groups and their end-release 
computations for the Lock and Dam 3 model and other similar structural 
systems, is given in Appendix A. Appendix I discusses the Rayleigh damping 
feature of the structural model, with section 1.2 giving information specific 
to the Lock and Dam 3 model. The numerical methods to be used in the 
solution of the equations of motion are either HHT-a or Wilson- 0 , which 
are discussed in Appendix F and G, respectively. 

2.7 Validation of lmpact_Deck Computer Program 

The validation of Impact_Deck computer program using the Lock and Dam 
3 model was made against the results obtained from the computer program 
SAP2000. The beam for validation has a total length of 838.666 ft long. In 
the validation procedure, the beam was modeled with 137 nodes and 136 
beam elements. The 7 inter-monolith pin connections (i.e., no bending 








ERDC/ITL TR-16-1 


33 


moment transfer between adjacent monoliths) were included in the model. 
A set of linear springs was located at the node where the pile supports were 
placed. 1 The strength of the concrete was assumed as fc = 5,000 psi with a 
corresponding modulus of elasticity for the concrete of E = 580,393.25 
kips/ft 2 . The beam cross-sectional area and the beam second moment of 
area (moment of inertia) were 110.0 ft 2 and 4436.666 ft 4 , respectively. The 
mass per linear foot of beam was calculated as m = 0.5127 kip *sec 2 /ft. A 
damping factor of 0.02 (i.e., 2% of the critical damping) was used in both 
computer program models. The impact-force time history was the Winfield 
test # 10 (generated using Impact_Force, Ebeling et al. 2010) and shown in 
Figure 2.30. The tangential-force time history was set equal to the 
transverse-force time history multiplied by a dynamic coefficient of friction 
of 0.5. The impact load was kept stationary at a point 402.896 ft along the 
impact deck due to restrictions in loading for SAP2000. The load-deflection 
relationship was assumed as the one presented in Figure C-4 of Ebeling et 
al. (2012). 

Figure 2.31 shows the results obtained for the node where the load was 
applied at 402.896 ft along the impact deck. The results are consistent 
between the two programs. 



1 The SAP2000 analysis is restricted to a linear spring model for each group of clustered piles. 



















ERDC/ITL TR-16-1 


34 


Figure 2.31 Validation of lmpact_Deck against SAP2000. 


Response at 402.89 feet 
Damping ratio = 2% 
-T 0 = 0.408 sec. 



— — lmpact_Deck-Linear-v = 0 ft/s — — SAP2000-Linear-v = Oft/s 


2.8 Numerical Example of the Elastic-Plastic Response Using 
lmpact_Deck 

In this section, results from a numerical example are shown that 
demonstrate the plastic behavior capability of the nonlinear impact deck 
clustered pile springs. Plastic response can develop if the limiting elastic 
displacement specified by the user (i.e., Point l in Figure 2.29) is low 
enough to force the springs to enter into the zone of plastic response. The 
input data for the Impact_Deck computer program for Lock and Dam 3 
model included the total length of the beam at 838.666 ft long with 137 
nodes and 136 beam elements. The 7 inter-monolith pin connections (i.e., 
no bending moment transfer between adjacent monoliths) were included in 
the model. A set of springs was located at the node where the pile supports 
were placed. The strength of the concrete was assumed as/c = 5,000 psi 
producing a modulus of elasticity for the concrete of E = 580,393.25 
kips/ft 2 . The beam cross-sectional area and the beam second moment of 
area (moment of inertia) were 110.0 ft 2 and 4436.666 ft 4 , respectively. The 
mass per linear foot of beam was calculated as m = 0.5127 kip *sec 2 /ft. A 
damping factor of 0.02 (i.e., 2% of the critical damping) was used. The 
impact-force time history applied was the Winfield test # 10 as shown in 
Figure 2.30. The tangential-force time history was set equal to the 
transverse-force time history but multiplied by a dynamic coefficient of 
friction of 0.5. In this example, the load was assumed to be in motion along 
the impact deck at a velocity of v = 3 ft/sec, starting at the node located at x 
= 402.896 ft. The spring stiffnesses assigned in this analysis did not make 





























ERDC/ITL TR-16-1 


35 


use of the Ebeling et al. (2012) Appendix A push-over analysis results for 
the 6-ft-diameter vertical piling because the Winfield Test # 10 loads could 
not be guaranteed to bring the computed stiffnesses into the zone of plastic 
deformation. In an effort to illustrate the effects of plastic deformation, the 
force-displacement relationship (backbone curve) was assumed to have the 
following slopes (stiffness), ki = 540.0 kip/ft, k2 = 215.0 kip/ft, k3 = 100.46 
kip/ft, and the stiffness when it is in the plastic unload path of kunioad = 
3-33*ki. The limit for the elastic displacement was assumed as Sdastic = 
0.0633 ft. The force-displacement relationship (backbone curve) had a 
second break point (second to third slope) at a displacement of 0.30 ft. That 
meant that the force value at the inflection point of the slope occured first at 
34.182 kips and a second inflection point at 85.073 kips. Figure 2.32 shows 
the results obtained for the node where the load was applied at 402.896 ft. 
The results in Figure 2.32 show the effect of exceeding the first yield point in 
the spring model. The purple values are offset from their original position 
by approximately 0.09 ft after the spring force reached its yield point. The 
red curve indicates the behavior for a linear elastic spring model where 
yielding does not occur. Figure 2.33 shows the plastic behavior of the spring 
at 402.896 ft, where the normal impact load had its initial point of contact. 
After 3.63 sec, the linear response oscillates around zero displacement and 
the plastic response oscillates around 0.098 ft. These behaviors can be 
observed in Figure 2.33 where the plastic response was reached (second 
slope in the force-displacement diagram), ending with a permanent 
displacement of around 0.098 ft. 

Figure 2.32 Dynamic response of the transverse spring located at x = 

402.896 ft. 


Response at 402.89 feet 
Damping ratio = 2% 
-T 0 = 0.408 sec. 



lmpact_Deck-Linear-v = 3 ft/s lmpact_Deck-Nonlinear-v= 3 ft/s 




















ERDC/ITL TR-16-1 


36 


Figure 2.33 Dynamic response of the transverse spring located at x = 

402.896 ft. 


Nonlinear Force-Displacement 
Elastic = 0.0633 ft : kl =540.0 kip/ft :k2 = 215.0 kip/ft : 
k3 = 100.46 kip/ft : k unload = 3.33*kj 



0 0.2 0.4 0.6 0.8 1 1.2 

Lateral Displacement (ft) 


2.9 lmpact_Deck GUI results 

In section 5, the visualization of data using the Impact_Deck GUI post¬ 
processing capabilities will be discussed, but these capabilities are being 
introduced here to give an idea of the output results from the 
Impact_Deck processing code, which used the formulation for impact deck 
structures discussed in section 2.8. The results are from the Lock and Dam 
3 example problem in section 5. The impact deck geometry and its 
material properties were the same as in section 2.8, with the exception of 
the load parameters (starting position and velocity along the approach 
wall) and the transverse non-linear spring properties of the individual pile 
cluster sub-system. The load still used the Winfield Test # 10 impact-force 
time history, but the starting position was moved to 501.188 ft from the 
beginning of the wall and had a velocity of 1.0 ft/sec (along the wall). 

The calculation of pile group stiffness was determined through push-over 
analyses performed on a single pile group. The Lock and Dam 3 impact- 
deck pile group consisted of a fixed head system of one vertical and two 
batter piles, all with a diameter of 24 in. In this case, the push-over 
analysis took all the piles into account, including the effects of batter 
(Figure 2.34). 

























ERDC/ITL TR-16-1 


37 


Figure 2.34 Reprint of the Figure 4-4 transverse direction of loading 
push-over analyses from Ebeling et al. (2012); fixed head results in 
green and pinned head results in blue. 



The push-over analysis for a transverse load on the fixed head, wet site 
analysis (dashed green curve shown in Figures 4-4 and B-8 on pages 81 
and 153, respectively, of Ebeling et al. 2012) was defined by the points 
listed in Table 2.1. These values were used to define the transverse spring 
model for an individual group of 3 piles. 

For the longitudinal spring model, the longitudinal forces acting on the 
impact deck when the two plastic hinge points occured were 1850 kips and 
2970 kips, respectively. These forces were much greater than and likely 
from a barge train impact event; thus no yielding of any pile groups in the 
longitudinal direction was anticipated. The data contained in Table 2.2 
were developed using the push-over analysis procedure outlined in section 
3 or Appendix A of Ebeling et al. (2012) for loading applied in the 
transverse direction. 



























ERDC/ITL TR-16-1 


38 


Table 2.1 Primary loading curve for the transverse spring model of a single pile group with a leading 
vertical pile followed by two batter piles (Lock and Dam 3). 


Force 

Deflection 

Notes on Flexural Plastic Hinging Conditions 

(kips) 

(inches) 

(feet) 

From Ebeling et al. (2012) 

0.0 

0.0 

0.0 


180.0 

2.462 

0.20517 

Pile to pile cap moment capacity reached 

250.0 

4.899 

0.40825 

Pile 3 yields in axial tension 

272.0 

7.544 

0.62867 

Flexural plastic hinges develop in piles below 
mudline 

317.75 

13.044 

1.087 

Pile 2 buckles 


Table 2.2 Primary loading curve for the longitudinal spring model of a single pile group with a 
leading vertical pile followed by two batter piles (Lock and Dam 3). 


Force 

Deflection 

Notes on Flexural Plastic Hinging Conditions 

(kips) 

(inches) 

(feet) 

0.0 

0.0 

0.0 


123.0 

4.18 

0.34833 

Pile to pile cap moment capacity reached 

149.7 

7.62 

0.635 

Flexural plastic hinges develop in piles below 
mudline 


According to Figure B-3 of Ebeling et al. (2012), yielding occurred for a 
pile when the moment exceeded 8544 kip-inch with no significant axial 
loading (i.e., pure bending). 

These Table 2.2 values come from the push-over analysis of the 3 pile 
system with 2 batter piles that were solved as a system, and therefore, the 
transverse and longitudinal push-over analyses must be performed using 
CPGA. 

The transverse force-displacement relationship (backbone curve) was 
therefore assumed to have the following slopes (stiffness), ki = 877.32 
kip/ft, k2 = 344.69 kip/ft, k3 = 99.81 kip/ft, and the stiffness when it is in 
the plastic unload path of kunioad = i.o*ki. The limit for the elastic 
displacement was assumed as deiastic = 0.20517ft. This backbone curve is 
shown in Figure 2.28a. This section does not provide an engineering 
analysis, but gives an idea of the information provided so that an 
engineering analysis might be made. 





ERDC/ITL TR-16-1 


39 


Nodal outputs provided from the FEO 1 analysis of an impact deck were the 
longitudinal displacement, transverse displacement, and rotational 
displacement (in radians) for each node at each time step of the simulation. 
A Table was also provided that gives the maximum displacements (longi¬ 
tudinal, transverse, and rotational) for each node and the time that the 
maximum displacement occurred. 


Figure 2.35 shows the GUI table of maximums for the example problem in 
this section. From this GUI table, it is possible to tell the time step and the 
node with the maximum displacement for transverse, longitudinal, and 
rotational displacements (as shown in Table 2.3). 


Figure 2.35 lmpact_Deck GUI Table of maximum nodal displacements 
for the L&D3 example impact deck. 



* File 

1 lmpact_Deck_TestProblem1.feo 


® Display extreme values and their times 

© Plot displacement vs. time Node Index 

| © Plot animated [MHIEIIIIHIH] 

[l t| Displace 

ment/Rotation: |x ▼! 

7] Loop Displacement/Rotation: 

(*Z3 

|0.00000 @| G 


Long. 

Long. 

Trans. 

Trains. 

Rot. 

Rot. 

- 

Node 

Disp. 

Time 

Disp. 

Time 

Disp. 

Time 


ID 

(ft) 

(sec) 

(ft) 

(sec) 

(ft) 

(sec) 


1 

0 

0 

0 

0 

0 

0.138 


2 

0 

0.138 

0 

0.21 

0 

0.138 


3 

0 

0.138 

0 

0.21 

0 

0.138 


4 

0 

0.138 

0.0001 

0.21 

0 

0.138 


5 

0 

0.138 

0.0001 

0.21 

0 

0.136 


6 

0 

0.138 

0.0001 

0.21 

0 

0.136 


7 

0 

0.136 

0.0001 

0.21 

0 

0.132 


8 

0 

0.136 

0.0002 

0.21 

0 

0.084 


9 

0 

0.136 

0.0002 

0.21 

0 

0.08 


10 

0 

0.134 

0.0002 

0.21 

0 

0.112 


11 

0 

0.134 

0.0002 

0.21 

0 

0.11 


12 

0 

0.132 

0.0003 

0.21 

0 

0.144 


13 

0 

0.13 

0.0003 

0.21 

0 

0.144 


14 

0 

0.082 

0.0003 

0.21 

0 

0.142 


15 

0 

0.08 

0.0003 

0.21 

0 

0.142 


16 

0 

0.078 

0.0004 

0.21 

0 

0.142 


17 

0 

0.076 

0.0004 

0.21 

0 

0.142 


18 

0 

0.076 

0.0004 

0.21 

0 

0 


19 

0 

0.076 

0.0004 

0.21 

0 

0.082 


20 

0 

0.108 

0.0005 

0.21 

0 

0.082 


21 

0 

0.14 

0.0005 

0.21 

0 

0.082 


22 

0 

0.134 

0.0005 

0.21 

0 

0.08 


23 

0 

0.13 

0.0005 

0.21 

0 

0.078 


24 

0 

0.086 

0.0006 

0.21 

0 

0.078 


25 

0 

0.084 

0.0006 

0.21 

0 

0.074 


26 

0 

0.082 

0.0006 

0.21 

0 

0.072 


27 

0 

0.08 

0.0006 

0.21 

0 

0.068 


28 

0 

0.078 

0.0007 

0.21 

0 

0.066 

- 


Table 2.3 Maximum nodal displacements for the lmpact_Deck example problem. 



Node number 

Value 

Time (seconds) 

Transverse 

86 

0.0663 feet 

0.26 

Longitudinal 

137 

0.002 feet 

0.212 

Rotational 

93 

0.0008 radians 

0.26 


1 FEO is the Finite Element Output format for the lmpact_Deck program. 































ERDC/ITL TR-16-1 


40 


Figures 2.36, 2.37, and 2.38 show the time histories for displacements at 
nodes 86,137, and 93, respectively. Because some of these displacements 
were very small, and the data were stored with limited precision, some of 
these plots developed ‘jaggies’, where the data seems to form stair steps. 


Figure 2.36 Transverse nodal displacement time histories for node 86. 














































ERDC/ITL TR-16-1 


41 


Figure 2.38 Rotational nodal displacement time histories for node 93. 



The Impact_Deck GUI also allowed the user to visualize the entire beam in 
motion using an exaggerated plot of displacements transversely, longitudi¬ 
nally, and rotationally. Figures 2.39, 2.40, and 2.41 show the displacements 
of the wall from these animated plots at the moment where the maximum 
displacement occurred, 0.26 sec for longitudinal displacements, 0.21 sec for 
transverse displacements, and 0.26 sec for rotational displacements. These 
data were also subject to the jaggies because of the precision of the stored 
data. The data were scaled to fit the plot. 

Element outputs provided from the FEO analysis of an impact deck were 
the axial force, shear force, and moment for each element at each time step 
of the simulation. A table was also provided that gives the minimum and 
maximum forces and moments for each element and the times that the 
minimum and maximum force/moment occured. 

Figures 2.42 - 2.44 show GUI tables of extreme values for the example 
problem in this section. From this GUI table, it is possible to tell the time 
step and the element with the extreme axial force, shear force, and 
moment (as shown in Table 2.4). 























ERDC/ITL TR-16-1 


42 


Figure 2.39 Transverse wall displacements at 0.26 sec. 



Figure 2.40 Longitudinal wall displacements at 0.21 sec. 


















































ERDC/ITL TR-16-1 


43 



Figure 2.42 lmpact_Deck GUI table of element minimum and maximum axial 
forces for the L&D3 example impact deck. 






























































ERDC/ITL TR-16-1 


44 


Figure 2.43 lmpact_Deck GUI table of element minimum and maximum shear 
forces for the L&D3 example impact deck. 



® Display extreme values and their times 
O Plot force/moment vs. time Element Index; [l -1 Mode: [/teat 


O Plot animated \i } || H 0 00000 \rj B Loop Mode: (tad 


135 

136 


10 

11 


13 

14 


16 

17 


19 

20 


22 

23 


- 1.8 

-0.45 

Shear 

Minimum 

Force 

(kip) 

-0.93 

-0.91 

- 0.86 

-0.76 

-0.64 

-0.5 

-0.34 

- 0.2 

-0.17 

-0.25 

-0.4 

-0.54 

- 0.66 

-0.75 

-0.82 

-0.87 

-0.91 

-0.94 

-0.99 

-1.04 

-1.05 

-1 

-0.9 

-0.76 


0.044 

0.042 

Shear 

Minimum 

Time 

(sec) 

0.142 

0.142 

0.142 

0.14 

0.14 

0.14 

0.138 

0.13 

0.118 

0.112 

0.144 

0.142 

0.142 

0.14 

0.138 

0.136 

0.134 

0.132 

0.13 

0.09 

0.09 

0.088 

0.086 

0.084 


1.8 

0.45 

Shear 

Maximum 

Force 

(kip) 

0.93 

0.91 

0.86 

0.76 

0.64 

0.5 

0.34 

0.2 

0.17 

0.25 

0.4 

0.54 

0.66 

0.75 

0.82 

0.87 

0.91 

0.94 

0.99 

1.04 

1.05 

1 

0.9 

0.76 


0.044 

0.042 

Shear 

Maximum 

Time 

(sec) 

0.142 

0.142 

0.142 

0.14 

0.14 

0.14 

0.138 

0.13 

0.118 

0.112 

0.144 

0.142 

0.142 

0.14 

0.138 

0.136 

0.134 

0.132 

0.13 

0.09 

0.09 

0.088 

0.086 

0.084 


Figure 2.44 lmpact_Deck GUI table of element minimum and maximum 
moments for the L&D3 example impact deck. 














































ERDC/ITL TR-16-1 


45 


Table 2.4 Extreme forces/moments for the lmpact_Deck example problem. 




Element number 

Value 

Time (seconds) 

Axial 

min. 

82 

-271.67 kips 

0.208 

max 

81 

528.44 kips 

0.200 

Shear 

min. 

82 

-293.15 kips 

0.202 

max 

81 

702.41 kips 

0.200 

Moment 

min. 

81 

-5,468.09 kip-feet 

0.180 

max 

82 

5,662.18 kip-feet 

0.202 


Because the simulated barge train impact occured between nodes shared 
by elements 81 and 82 , the extreme forces were found in those elements. 
Figures 2.45 - 2.50 show the time histories for the axial force, shear force, 
and moments for the elements with the minimum and maximum values, 
respectively. 

Figures 2.51 - 2.53 show the axial force, shear force, and moment at 0 . 200 , 
0 . 200 , and at 0.180 sec, respectively, for the entire wall. These forces and 
moments were collected for animated visualization in the Impact_Deck 
GUI. 


Figure 2.45 Axial-force time histories for element 82. 











































ERDC/ITL TR-16-1 


46 


Figure 2.46 Axial-force time histories for element 81. 


FEO Element Output 


• Impact JDeck_TestProbleni1 .feo 


© Display extreme values and their times 
© Plot force/moment vs. time Element Index [si 
Plot animated ^ || ^ M G.C3303 


] Mode: | Axial ▼] 

^j| 0 Loop Mode: [Axial 


Y 

L 


Axial Force 
























































ERDC/ITL TR-16-1 


47 



Figure 2.49 Moment time histories for element 81. 


FEO Element Output 


lmpact_Deck_Test Problem 1 _pushover.feo 


© Display extreme values and their times 

® Plot fbrce/moment vs. time Element Index [81 Mode: [ Moment -• | 

Plot animated ^ || ^ C.3C333 


Y 

L 


Moment (kip-ft) 


moment (kip 




















































ERDC/ITL TR-16-1 


48 


Figure 2.50 Moment time histories for element 82. 






File 

: lmpact_Deck_Test Problem 1 _pusho ver.f eo 

1 © Display extreme values and their times 




1 (®) Plot force/moment vs. time Element Index: [82 

▼ | Mode: |Moment 



Plot animated ^ || ^ M ; 

[4JI13 Loop Mode: |Axial 

13 



Y 

L 


Elem = 82 


Moment (kip-ft) 



Figure 2.51 Wall axial forces at 0.2 sec. 

















































ERDC/ITL TR-16-1 


49 



Figure 2.53 Wall moments at 0.18 sec. 



Of primary importance for doing a pile group founded flexible wall analysis, 
is being able to see exactly how much force each pile group will be able to 
resist during an impact event. This is measured by finding the resisting 
force from the spring model used for the pile group for the pile group’s 


















































ERDC/ITL TR-16-1 


50 


displacement. Figure 2.54 shows the table of maximum forces/moments 
resisted at the nodes representing a pile group in the impact deck example 
model. Looking through this table gives the information that the maximum 
individual pile group response force is in the transverse direction at node 85 
and at time 0.26 sec. This peak response force is 54.465 kips. Figure 2.55 is 
the same window scrolled down to reveal the maximum displacements of 
these pile group nodes, and when these displacements occurred. 

A second table in the program reveals the response forces of each pile 
group node at a specified time. Figure 2.56 shows this table for time 
0.26 sec. At the bottom of this table, the individual responses are summed 
to reveal the total force in each direction. Information about the impulse 
calculations over all time is also given at the end of this table. For time 
0.26 sec, the total force from the 96 pile groups acting in the transverse 
direction is 722.776 kips. For the entire run, the transverse impulse was 
454.31 kip-sec. 

Figure 2.57 shows the response of the pile group at node 85 at time 
0.260 sec as a force-versus-displacement plot, and as force and 
displacement time histories. This is the time when the displacement for 
the pile group at node 85 is at its maximum location in the transverse 
direction. In this example, the longitudinal and rotational displacement 
created small resisting forces (Figures 2.58 and 2.59). 


Figure 2.54 Table of pile group response maximum forces and moments and their 

time. 


Pile Group Response 






File 







lmpact_Deck_TestProblem1.feo 






0 Display extreme values and their times 





O Display Spring Forces at Time: 0.0000 





- Plot Animated IMIIHII ► IIIIIIMIIMI JoSOQOO 

[3 Loop Pile Group 

Node Index [ 1 

▼ | Degree of Freedom: |x ▼ ! 


Long. 

Long. 

Trans. 

Trans. 



* 

Node 

Force 

Time 

Force 

Time 

Moment 

Time 


ID 

(kips) 

(sec) 

(kips) 

(sec) 

(kip-ft) 

(sec) 


1 

251.4489 

0.21 

0.9333 

0.142 

0.0012 

0.0012 

u 

2 

0.0047 

0.21 

0.001 

0.18 

0 

0 


3 

0.0138 

0.21 

0.0028 

0.18 

0 

0 


4 

0.0229 

0.21 

0.0045 

0.18 

0 

0 


5 

0.032 

0.21 

0.006 

0.18 

0 

0 


6 

0.0411 

0.21 

0.0073 

0.178 

0 

0 


7 

0.0502 

0.21 

0.0083 

0.178 

0 

0 


8 

0.0593 

0.21 

0.009 

0.178 

0 

0 


9 

0.0684 

0.21 

0.0093 

0.178 

0 

0 


10 

0.0775 

0.21 

0.0093 

0.178 

0 

0 


11 

0.0866 

0.21 

0.0089 

0.176 

0 

0 


12 

0.0957 

0.21 

0.0082 

0.176 

0 

0 


13 

0.1048 

0.21 

0.0078 

0.086 

0 

0 


14 

0.1139 

0.21 

0.0082 

0.082 

0 

0 


15 

0.123 

0.21 

0.0088 

0.08 

0 

0 


16 

0.1321 

0.21 

0.0096 

0.078 

0 

0 


17 

0.1412 

0.21 

0.0105 

0.076 

0 

0 


19 

0.1505 

0.21 

0.0092 

0.076 

0 

0 


20 

0.1596 

0.21 

0.0059 

0.072 

0 

0 


21 

0.1687 

0.21 

0.0059 

0.14 

0 

0 


22 

0.1778 

0.21 

0.007 

0.134 

0 

0 


23 

0.1869 

0.21 

0.0087 

0.13 

0 

0 


24 

0.196 

0.21 

0.0105 

0.126 

0 

0 


25 

0.2051 

0.21 

0.0121 

0.124 

0 

0 


26 

0.2143 

0.21 

0.0134 

0.124 

0 

0 


27 

0.2234 

0.21 

0.0144 

0.122 

0 

0 


28 

0.2325 

0.21 

0.0148 

0.12 

0 

0 


29 

0.2416 

0.21 

0.0149 

0.118 

0 

0 

- 



















ERDC/ITL TR-16-1 


51 


Figure 2.55 Table of pile group response maximum displacements and their time. 


lmpact_Deck_TestProbleml .feo 


. o Display extreme values and their times 

1 O Display Spring Forces at Time: 0.0000 



O Plot Animated M M ► II N 0.00000 

0 Loop Pile Group Node Index [l 

▼ | Degree of Freedom: [x 


134 

0.7024 

0.212 

0.0669 

0.214 

0 

0 

135 

0.7024 

0.212 

0.0247 

0.214 

0 

0 

136 

0.7024 

0.212 

0.0029 

0.212 

0 

0 


Long. 

Long. 

Trans. 

Trans. 



Node 

Disp. 

Time 

Disp. 

Time 

Rot 

Time 

ID 

(in) 

(sec) 

(in) 

(sec) 

(rad) 

(sec) 

1 

IE-05 

0.08 

0 

0 

0 

0 

2 

4E-05 

0.162 

0 

0 

0 

0 

3 

6E-05 

0.152 

IE-05 

0.178 

0 

0 

4 

9E-05 

0.172 

IE-05 

0.094 

0 

0 

5 

0.00012 

0.196 

IE-05 

0.09 

0 

0 

6 

0.00014 

0.176 

IE-05 

0.086 

0 

0 

7 

0.00017 

0.192 

IE-05 

0.084 

0 

0 

8 

0.00019 

0.178 

IE-05 

0.082 

0 

0 

9 

0.00022 

0.19 

IE-05 

0.08 

0 

0 

10 

0.00025 

0.208 

IE-05 

0.076 

0 

0 

11 

0.00027 

0.19 

IE-05 

0.074 

0 

0 

12 

0.0003 

0.202 

IE-05 

0.07 

0 

0 

13 

0.00032 

0.188 

IE-05 

0.068 

0 

0 

14 

0.00035 

0.198 

IE-05 

0.064 

0 

0 

15 

0.00037 

0.188 

IE-05 

0.062 

0 

0 

16 

0.0004 

0.196 

IE-05 

0.06 

0 

0 

17 

0.00041 

0.19 

IE-05 

0.058 

0 

0 

19 

0.00045 

0.194 

IE-05 

0.062 

0 

0 

20 

0.00048 

0.2 

IE-05 

0.13 

0 

0 

21 

0.0005 

0.192 

IE-05 

0.122 

0 

0 

22 

0.00053 

0.198 

IE-05 

0.116 

0 

0 

23 

0.00056 

0.21 

IE-05 

0.112 

0 

0 

24 

0.00058 

0.196 

IE-05 

0.11 

0 

0 


Figure 2.56 Table of pile group responses for each pile group individually and 
summed (not shown) at time 0.26 sec. 


Pile Group Response 







File 

| lmpact_Deck_TestProbleml.feo 

|| © Display extreme values and their times 






II ® Display Spring Forces at Time: jl2600 

IT 





|l © Plot Animated M M ► II N HI 100000 

[£-J| 0 Loop Pile Group Node Index [l 

▼ | Degree of Freedom: [x 

d 










[Time: 

0.26 






n 


Long. 

Trans. 






Node 

Force 

Force 

Moment 




ID 

(kips) 

(kips) 

(kip- 

ft) 



M 

1 

- 0.302 

0 


0 




2 

0.0041 

0.0003 


0 




3 

0.0121 

0.0008 


0 




4 

0.0201 

0.0013 


0 




5 

0.0281 

0.0018 


0 




6 

0.0361 

0.0021 


0 




7 

0.0442 

0.0024 


0 




8 

0.0522 

0.0025 


0 




9 

0.0602 

0.0024 


0 




10 

0.0682 

0.0022 


0 




11 

0.0762 

0.0019 


0 




12 

0.0842 

0.0014 


0 




13 

0.0922 

0.0008 


0 




14 

0.1002 

0.0001 


0 




15 

0.1082 

- 0.0007 


0 




16 

0.1163 

- 0.0016 


0 




17 

0.1243 

- 0.0025 


0 




19 

0.1325 

- 0.0029 


0 




20 

0.1405 

- 0.0027 


0 




21 

0.1485 

- 0.0025 


0 




22 

0.1565 

- 0.0022 


0 




23 

0.1645 

- 0.0018 


0 




24 

0.1725 

- 0.0013 


0 




25 

0.1805 

- 0.0006 


0 




26 

0.1886 

0.0002 


0 




27 

0.1966 

0.0011 


0 



- 











































ERDC/ITL TR-16-1 


52 


Figure 2.57 Transverse pile group responses for the pile group at node 85 and at 

time 0.24 sec. 



Figure 2.58 Longitudinal pile group responses for the pile group at node 85 and 

at time 0.24 sec. 


File 

lmpact_Deck_TestProbleml feo 

© Display extreme values and their times 
© Display Spring Forces at Time: [0.2600 

® Plot Animated [MlfflfFlffllfflBl] [026000 fjj| m Loop Pile Group Node Index [85 Degree of Freedom: [y 

Time = 0.26000 Node: 85 Y-Axis 


X 


Y 


u 















































ERDC/ITL TR-16-1 


53 


Figure 2.59 Rotational pile group response for the pile group at node 85 and at 

time 0.24 sec. 



The Case for Dynamic Analysis: 

This example problem demonstrated the need to perform a dynamic 
analysis of these pile founded walls. This example used an impact time 
history that was the result of the Winfield Test # 10. This impact time 
history was applied transverse to the approach wall at a position starting 
at 501.187 ft along the wall (at the same location as node 82) and moved at 
1 ft/sec along the approach wall. The peak force for the impact time history 
was 516.4 kips at time 0.174 sec. 

Node 85 (as discussed above) had a maximum transverse force of 56.4652 
kips, which occurred at 0.26 sec. This was the last pile group node along a 
monolith, so its transverse displacement was affected by the moment 
induced (to the monolith it is connected) by the impact load. The total pile 
group response for the whole impact deck structure (96 pile group nodes) 
at time 0.26 sec was 722.775 kips, which was greater than the peak input 
force. It was also the peak overall response. 

At time 0.174, the impact load reached its first peak with a force of 
516.2 kips. Data was also collected at this time. Table 2.5 shows the values 
generated by this data collection. Figure 2.60 takes this information a step 
further by displaying the time histories for the three forces. 






























ERDC/ITL TR-16-1 


54 


Table 2.5 Transverse forces with respect to time. 


Time (s) 

Node 85 Force (kips) 

Impact Load Force 
(kips) 

Total Pile Group 
Response (kips) 

0.174 

39.131 

516.200 

324.254 

0.200 

47.277 

516.400 

611.247 

0.260 

56.465 

446.358 

722.776 


Figure 2.60 Time-history plot of transverse input forces, total force response for 
all the pile groups, and an individual pile group response forces. 



From Table 2.5 and Figure 2.60, it can be seen that the overall pile group 
response is dynamic because it does not track with the input force. 

Instead, the pile groups respond to the input force overtime due to inertial 
effects. Because of this, the overall pile group response can have higher 
peak forces. For example, at a time of peak response of 0.2 sec, a 
maximum impact force of 516.4 kips is applied to the approach wall. The 
total transverse-force response of 96 nodes represents all of the pile group 
and totals 611.25 kips. This 18 % larger force response is due to the 
contribution of the first two terms of the equation of motion (also seen in 
Appendix A) for the dynamic structural response: 


M{ii(t)} + [C]{u(t)} + [Jf]{u (0} = {-*=■(»)} 


(2.1) 






























ERDC/ITL TR-16-1 


55 


At time 0.26 sec, when a peak response force is recorded at node 85 for an 
individual pile group, the contribution of the first two terms of the 
equation of motion for the dynamic structural response is even larger 
resulting in an overall response force of 722.776 kips. This is nearly 40% 
greater than the peak input force of 516.4 kips. And an overall peak 
response force of 722.776 kips is 55% larger than the 446.358 kips input 
force imposed at 0.26 sec. These observations demonstrate the importance 
of applying the equation of motion for calculating pile group structural 
response forces (and displacements). These differences explain why a 
dynamic analysis is required versus a static analysis, which the user 
provided impact load is applied as a single peak value (e.g., determined to 
be the input peak force from the time history). 

Load Sharing in Dynamic Analysis: 

Another point to note is the individual pile group response has much lower 
peak forces than the input load. This is because the load is being shared 
with other pile groups along the 96 groups that support the eight impact 
deck monoliths, which sum up to the overall pile group response. The peak 
response of node 85, at 56.465 kips, is only 11% of the peak input force of 
516.4 kips and 8% (1/12) of the overall peak response force of 722.776 kips. 
Because there are 16 pile group nodes per monolith, it may be surmised 
that the peak individual response force would be 1/ 16 th instead of 1/ 12 th of 
the overall response. Note the fact that this peak force is obtained only at 
node 85 and includes transverse displacement due to rotational moment 
and inertia, which are difficult to predict without dynamic analysis. 

Validation Using Impulse Calculations: 

The higher peak values of the overall pile group response seem out of place 
until an impulse calculation (taking the area beneath the time history 
curves for input load and overall pile group response in Figure 2.60) is 
performed. Despite inertial effects, the impulse of the input load must be 
equivalent to the overall pile group response, if the piles do not fail. When 
an impulse calculation is performed for the overall pile group response, 
the result is 454 kip-sec. The impulse for the input force is 463 kip-sec. 

The difference is minimal (<2%) and easily explained by the fact that these 
are only transverse forces and do not include moment arm effects that are 
induced by the impact load on the impacted monolith. 



ERDC/ITL TR-16-1 


56 


2.10 Final Remarks 

In this section, the Lock and Dam 3 physical model was presented and the 
mathematical model to calculate the dynamic response was also 
developed. Impact_Deck, a computer program, was used to calculate the 
dynamic response of an elastic beam supported over linear elastic or 
plastic spring supports. The mathematical formulation featured a method 
to calculate the end release at the inter-monolith connection. The impact 
normal and parallel concentrated external load was located at a specified 
location or assumed to have motion at a specified constant velocity. The 
damping effect was considered by means of the Rayleigh damping model, 
which depended on the natural frequencies of the system. These natural 
frequencies were calculated in an approximate way by using the linear 
stiffness of the springs and the mass per unit length of the beam. The 
results of Impact_Deck proved to be valid when compared to the results 
obtained with SAP2000. An example was presented to show the plastic 
behavior of the springs and how these results compared to the linear 
elastic response. 



ERDC/ITL TR-16-1 


57 


3 An Approach Wall with Impact Beams on 
Nontraditional Pile Supported Bents - 
McAlpine Example 

3.1 Introduction 

This chapter summarizes an engineering methodology and corresponding 
software feature contained within the Impact_Deck software for performing 
a dynamic structural response analysis of a flexible impact beam supported 
over groups of clustered vertical piles configured in a triangle pattern and 
subjected to a barge impact event. This example is based on an alternative 
approach wall design proposed for the McAlpine Locks and Dam on the 
Ohio River in Kentucky. 

This example models a glancing blow impact event of a barge train 
impacting an approach wall as it aligns itself with a lock is an event of short 
duration; the contact time between the impact corner of the barge train and 
the approach wall can be as short as a second or as long as several seconds. 
The next generation of Corps approach walls is more flexible than the 
massive, stiff-to-rigid structures constructed in the past in order to reduce 
construction costs as well as to reduce damage to barges during glancing 
blow impacts with lock approach walls. A flexible approach wall or flexible 
approach wall system is one in which the wall has the capacity to absorb 
impact energy by deflecting or “flexing” during impact, thereby affecting the 
dynamic impact forces that develop during the impact event. 

3.2 Alternative Flexible Approach Wall - Physical Model 

McAlpine alternative flexible approach wall is located in Louisville, 
Kentucky, at the Falls of the Ohio. McAlpine Locks and Dam are located at 
mile point 606.8 and control a 72.9-mile-long (117.3 km) navigation pool. 
The McAlpine locks underwent a 10-year expansion project that was 
completed in early 2009. The flexible approach wall system discussed in 
this section was an alternative flexible approach wall design. 

An artist rendering of the McAlpine alternative flexible approach wall (plan 
and cross-sectional views) is shown in Figure 3.1. The structure consists of 
continuous, flexible concrete beams with a span of approximately 96 ft per 
segment. Each end of the beam rests on a pile cap that is supported by a 



ERDC/ITL TR-16-1 


58 


group of three piles. The supporting piles provide a flexible resistance to the 
barge train impact forces that are applied to the beam. The continuity of the 
impact beam with the pile cap impact feature is achieved by means of shear 
key at each pile group support. That means that the ends of each beam 
transfers the longitudinal and transverse forces with no moment transfer to 
each pile group support (i.e., a pinned support). The axial and transverse 
forces at the end of the beams are transferred to the pile cap by means of a 
shear key. The shear key portion of the pile cap structural feature has a 
length of 11.5 ft. The length of the shear key is the distance between two 
consecutive concrete beams in this figure. The massive pile cap rests over a 
clustered group of three vertical piles. Each pile is 5 ft 8 in. in diameter. 
They are arranged in a triangular pattern so as to absorb the torsion 
generated at the pile group due to the eccentricity between the center of the 
pile group and the location of the end of each of the flexible impact beams 
that are supported by a clustered pile group. 


Figure 3.1 McAlpine alternative flexible approach wall. 



3.3 McAlpine Alternative Flexible Approach Wall - Mathematical 
Model 

The McAlpine alternative flexible approach wall can be modeled using beam 
elements because its length is much greater than its other two dimensions. 
The length of each flexible impact beam segment is approximately 96 ft and 
the width and height are 6 ft 7 in. and 9 ft, respectively. The mathematical 
model can be seen in Figure 3.2. In the analysis, the model consists of two 
consecutive beams with longitudinal and transverse elastic-plastic spring 
































ERDC/ITL TR-16-1 


59 


supports at the start and end of the system. These two set of springs model 
the pile group at the start and end of the two consecutive beams. The 
transverse non-linear spring represents the effect of the clustered group of 
three vertical piles. The longitudinal end spring represents the effect of the 
response of the end support pile group and that of the other groups of piles 
beyond this location. At the center pile group, three rigid links are used to 
model the high stiffness and mass of the pile cap. The nodes that connect 
the impact beams to the pile cap do not transfer moment between these 
impact beams and the center pile cap. The center pile group is modeled with 
three springs (two translational and one rotational). The translational 
spring stiffness is calculated by means of a push-over analysis (section 3 or 
Appendix A in Ebeling et al. 2012) and the rotational spring stiffness is 
calculated and seen in Appendix E. 


Figure 3.2 McAlpine flexible approach wall mathematical model. 



The mathematical model is formulated using 3-D beam elements. A 3-D 
beam element has 6 degrees of freedom per node, producing 12 degrees of 
freedom per element. The degrees of freedom per node are 3 translations 
and 3 rotations as shown in Figure 3.3. The force F x (t) applied normal to 
the flexible beams is the impact-force time history developed using the PC- 
based software Impact_Force (Ebeling et al. 2010). The force F y (t) is 
applied parallel to the wall and is a fraction of the normal force calculated 
using the dynamic coefficient of friction between the barge and the flexible 
beam contact surface in the Impact_Deck software. 

The model used to describe the beam is developed in the plane so the 
beam element has 3 degrees of freedom per node and 6 degrees of freedom 
per element. The degrees of freedom per node are 2 translations and 1 



























































ERDC/ITL TR-16-1 


60 


rotation (Figure 3.4). Based on the notation of Figure 3.3, the force and 
moment conditions for node i are Fi, x = Vi, Mi, x = o, Fi, y = Fi, Mi, y = o, Fi, z = 
o, and Mi, z = Mi, and for node/ are Ff x = Vf, Mf, x = o, Ff, y = Ff, Mf, y = o, Ff, z = 
o, and Mf z = Mf. To transform a 3-D beam element to a 2-D (plane 
element), the moment about the “x” axis, the moment about the “y” axis, 
and the force in the “z” directions are equal to zero. 


Figure 3.3 (a) Typical 3-D segment of the Impact Deck beam element, (b) Impact force applied 
to the Impact Deck, (c) Typical 3-D beam element. 



Figure 3.4 Typical 2-D beam element used in lmpact_Deck. 


V 

e u M t 

U j, 

Node i 



a 


f w fi V f 



Nodef u _i F f 


A push-over type of analysis was conducted to characterize the force- 
versus-displacement behavior of a cluster of three vertical piles under 
static lateral loading and thus define the stiffness coefficient of the two 
translational springs and one rotational spring. Appendix E summarizes 
the three-spring stiffness for the McAlpine alternative flexible wall. 

3.4 Nonlinear force-deflection relationship for the springs supports 

Section 2.5 discussed how the nonlinear spring response force- 
displacement backbone curve was used to model plastic deformation in 
the pile substructure for dynamically loaded structures. Push-over 




































ERDC/ITL TR-16-1 


61 


calculations were similarly performed for the pile layout for the McAlpine 
alternative flexible wall supports. 

The McAlpine alternative flexible wall bent was different from the other 
bents, in that its piles were not in a straight line. Its drilled-in-place (DIP) 
piles were placed in a triangle (per the following figure), and it was 
assumed that a push-over analysis would yield different results than for a 
traditional pile bent. 


Figure 3.5 Plan view of the flexible wall pile layout. 



After a push-over analysis in CPGA was performed, it was found that this 
was not the case, other than the fact that due to the size of the beams and 
the method of anchoring the beams to the pile cap superstructure the 
bents actually developed bending failure at the cap and mudline 
simultaneously. 

The transverse and longitudinal push-over, force-versus-displacement 
curves were effectively the same as taking an individual pile curve and 
multiplying the force by three. Tables 3.1 and 3.2 summarize the primary 
loading curves used to define the transverse and longitudinal spring 
models, respectively. 

Appendix E provides a method for calculating the rotational force-versus- 
displacement curve from the force-displacement curve for a single pile. 
The single pile curve is provided in Table 3.3. 




























ERDC/ITL TR-16-1 


62 


Table 3.1 Primary loading curve for the transverse spring model for a McAlpine alternative flexible wall 

bent (3 piles). 


Force 

Deflection 

Notes 

(kips) 

(inches) 

(feet) 

Adapted from Appendix A of Ebeling et al. 
(2012) 

0.0 

0.0 

0.0 


2141.5 

1.8 

0.15 

Pile to pile cap moment capacity reached 
at the same time hinge develops at 
equivalent depth of fixity 

2141.5 

4.0 

0.33 

Plastic hinge rotation 


Table 3.2 Primary loading curve for the longitudinal spring model for a McAlpine alternative flexible 

wall bent (3 piles). 


Force 

Deflection 

Notes 

(kips) 

(inches) 

(feet) 

Adapted from Appendix A of Ebeling et 
al. (2012) 

0.0 

0.0 

0.0 


2141.5 

1.8 

0.15 

Pile to pile cap moment capacity 
reached at the same time hinge 
develops at equivalent depth of fixity 

2141.5 

4.0 

0.33 

Plastic hinge rotation 


Table 3.3 Primary loading curve for a spring model for a single 6-ft diameter DIP pile. 


Force 

Deflection 

Notes 

(kips) 

(inches) 

(feet) 

Adapted from Appendix A of Ebeling et 
al. (2012) 

0.0 

0.0 

0.0 


713.8 

1.8 

0.15 

Pile-to-pile cap moment capacity 
reached at the same time hinge 
develops at equivalent depth of fixity 

713.8 

4.0 

0.33 

Plastic hinge rotation 


3.5 Solving for the motion of the structure 

The equations of motion for a flexible approach wall structure comprised 
of decks supported on clustered pile groups and their end-release 
computations for the McAlpine alternative flexible approach wall model 
and other similar structural systems are given in Appendix B. Appendix I 
discusses the Rayleigh damping feature of the structural model, with 
section I.3 giving information specific to the McAlpine alternative flexible 
approach wall model. The numerical methods to be used in the solution of 
the equations of motion are either HHT-a or Wilson- 0 , which are 
discussed in Appendix F and G, respectively. 






ERDC/ITL TR-16-1 


63 


3.6 Validation of lmpact_Deck Computer Program 

The validation of Impact_Deck computer program for the McAlpine 
alternative flexible approach wall model was made against the results 
obtained from the computer program SAP2000. The beam had a total 
length of 180.5 ft long. In the validation procedure, the beam was modeled 
with 24 nodes and 24 beam elements. The three rigid elements that model 
the pile cap support were included in the model. A set of linear elastic 
springs was located at node 1 and 23 where the start and end pile supports 
were placed. 1 The strength of the concrete was assumed as fc = 5,000 psi 
with a corresponding modulus of elasticity for the concrete of E = 
580,393.25 ksf. The beam cross-sectional area and the beam second 
moment of area (moment of inertia) were 54.668 ft 2 and 517.2 ft 4 , 
respectively. The mass per linear foot of beam was calculated as in = 

0.25486 kip *sec 2 /ft. A damping factor of 0.02 (i.e., 2% of the critical 
damping) was used in both computer program models. Figure 3.6, impact- 
force time history was the Winfield test # 10 (generated using Impact_ 
Force, Ebeling et al. 2010) and applied at node 11 (i.e., x = 84.5 ft). The 
tangential-force time history was set equal to the transverse time history 
multiplied by a dynamic coefficient of friction of 0.5. The impact load was 
kept stationary at node 11 due to restrictions in loading for SAP2000. A pile 
group rotational spring stiffness value of 12,426,909 kip*ft/rad was com¬ 
puted as outlined in the Appendix E calculation steps. This calculation used 
a stiffness value of 132,189.5 kip/ft as the initial segment of the push-over 
curve which was applied to each of the two translational springs for each 
pile of the three-pile groups. The stiffness of the three (two translational and 
one rotational) representing the central pile group was assigned to node 12’, 
located 7.708 ft in the transverse direction behind node 12. This position 
was located at the center of rotation of the pile group. Figures 3.7 to 3.10 
show the dynamic response time histories of nodes 1,12’, and 23. With the 
exception of a few minor differences, both computer programs resulted in 
the computation of essentially the same system response values. At nodes 1 
and 23, some differences in the magnitude of the computed transverse dis¬ 
placement values were apparent. However, the magnitudes of these values 
were very small, which was associated with numerical approximations. 
Figure 3.10 shows the rotation time histories computed at nodes 12 and 12’. 
These two nodes defined the rigid beam element of the pile cap, which was 
perpendicular to the impact beam model. The rotations were the same for 
both nodes and indicated a rigid element behavior. 


1 The SAP2000 analysis is restricted to a linear spring model for each group of clustered piles. 



ERDC/ITL TR-16-1 


64 


Figure 3.6 Force time history of Winfield Test # 10. 



- Normal Force - Parallel Force 


Figure 3.7 Validation of lmpact_Deck against SAP2000 - Transverse 
displacement at node 1. 

ImpactDeck - Linear - Linear - Elastic - v = 0 ft/s 
Flexible Wall - X - Direction 
(No moving load applied at node 11) 

(kr)l = 12,426,909 : (kr)2 = 12,426,909 



Time (sec) 

-ImpactDeck-Node 1 • • • • SAP2000-Node 1 























































ERDC/ITL TR-16-1 


65 


Figure 3.8 Validation of lmpact_Deck against SAP2000 - Transverse 
displacement at node 23. 

ImpactDeck - Linear - Elastic - v = 0 ft/s 
Flexible Wall - X - Direction 
(Mo moving load applied at node 11) 

(kr)l = 12,426,909 : (kr)2 = 12,426,909 



Time (sec) 

-Impact_Deck-Node 23 -SAP2000-Node 23 


Figure 3.9 Validation of lmpact_Deck against SAP2000 - Transverse 
displacement at node 12’. 


Impact_Deck 

Flexible Wall - X - Direction 
(kr)l = 12,426,909 : (kr)2 = 12,426,909 



Time (sec) 

- Impact_Deck-Node 12' - v=0 ft/s - Linear - SAP2000 -Node 12 - v=0 ft/s - Linear 

— SAP2000-\ode 12' - v=0 ft/s - Linear 



























































ERDC/ITL TR-16-1 


66 


Figure 3.10 Validation of lmpact_Deck against SAP2000 - Rotation at 

node 12 and 12’. 


Flexible Wall 

Node 12 - 12' - Rotation - Z 

(Load applied at node 11 moving at the specified velocity "v") 
(kr)l = 12,426,909 : (kr)2 = 12,426,909 



Time (sec) 

- Impact_Beam-Node 12 - v=0 ft/s - Linear • • • • Impact_Deck-Node 12' - v=0 ft/s - Linear 

- SAP2000-Node 12 - v=0 ft/s - Linear - SAP2000-Node 12' - v=0 ft/s - Linear 


3.7 Numerical Example of the Elastic-Plastic Response Using 
lmpact_Deck 

This section presents the results of a numerical example that demonstrates 
the plastic behavior capability of the nonlinear impact-deck clustered pile 
springs. Plastic response can develop if the limiting elastic displacement 
specified by the user (i.e., Point l in Figure 3.5) is low enough to force the 
springs to enter into the zone of plastic response. The input data for the 
Impact_Deck computer program for McAlpine alternative flexible approach 
wall were as follows: The beam has a total length of 180.5 ft long with 24 
nodes and 24 beam elements. The three rigid elements that model the pile 
cap support was included in the model. A set of linear elastic springs were 
located at node 1 and 23 where the start and end pile supports were placed. 
The strength of the concrete was assumed as fc = 5,000 psi producing a 
modulus of elasticity for the concrete of E = 580,393.25 ksf The beam 
cross-sectional area and the beam second moment of area (moment of 
inertia) were 54.668 ft 2 and 517.2 ft 4 , respectively. The mass per linear foot 
of beam was calculated as in = 0.25486 kip *sec 2 / ft. A damping factor of 
0.02 (i.e., 2% of the critical damping) was used. The force time history was 
the Winfield test # 10 and shown in Figure 3.6. The tangential-force time 
history was set equal to the transverse force time history but multiplied by a 
dynamic coefficient of friction of 0.5. In this example, the load was assumed 
to be in motion along the impact beam at a velocity of v = 3 ft/sec starting at 
the node located at x = 84.5 ft in one set of calculations and applied at x = 
84.5 ft along the beam for the entire duration of the impact event in the 

































ERDC/ITL TR-16-1 


67 


second set of calculations (i.e., with at v = oft/sec). In order to ensure 
plastic deformations, the spring models obtained by the push-over analysis 
were not used. Instead, the initial slopes for the stiffness of the two 
translational springs were assigned values equal to fa = 132,189.5 kip/ft and 
k 2 = 66,094.75 kip/ft with a stiffness for unload after the elastic 
displacement equal to kunioad = ki. The elastic displacement that defines the 
point of demarcation for elastic and plastic zone behavior was deiastw = 0.003 
ft. The rotational spring stiffness was set equal to far = 12,426,909.0 
kip*ft/rad and far = 6,213,454.5 kip*ft/rad with a stiffness for unload after 
the elastic displacement was equal to ( fanioadf = far. The elastic rotation that 
defines the point of demarcation for elastic and plastic zone behavior was 
Qelastic = 0.25 rad. The stiffness of the three (two translational and one 
rotational) representing the central pile group was assigned to node 12’, 
located 7.708 ft in the transverse direction behind node 12. Figure 3.11 
shows the dynamic response time history obtained for node 12 and 12’ (i.e., 
center of rigidity of the three pile group). The green nodal displacement 
trace in this figure shows a permanent lateral displacement of approxi¬ 
mately 0.0011 ft after the transverse pile group spring develops plastic 
behavior. Figure 3.12 shows the rotational behavior of node 12 and 12’. If the 
impact load is moving, the response shows a change in sign for the rotation 
indicating that the load moves from one rigid element to the next rigid 
element, that is, the load moves from the right element from the centerline 
of the pile cap to the left element. After 3.63 sec, the linear response 
oscillates around zero displacement and the plastic response oscillates 
around 0.0011 ft. The explanation for the noted behaviors is explained by 
the fact that Figure 3.13 shows that plastic response is reached (i.e., 
response along the second slope of the force-displacement diagram) ending 
with a permanent displacement of around 0.0011 ft. 

3.8 lmpact_Deck GUI results 

The Impact_Deck GUI was also used to run the McAlpine flexible wall 
problem in this section. This section does not provide an engineering 
analysis, but gives an idea of what information is provided so that an 
engineering analysis might be made. 

The inputs are the same as those used for validating the model, with the 
minor exception that the longitudinal velocity of the barge train was 1 ft/sec, 
and that the push-over analysis spring models for the pile groups were used. 
These changes did not allow the pile groups to go into plastic deformation 
during loading. 



ERDC/ITL TR-16-1 


68 


Figure 3.11 Dynamic transverse response of node 12 and 12’. 


ImpactDeck 

Flexible Wall - X - Direction 

(Load applied at node 11 moving at the specified velocity "v") 
(kr)l = 12.426.909 : (kr)2 = 12,426,909 
(kr)l = 12,426,909 : (kr)2 = 6213.454.52 



Time (sec) 

- Impact_Deck-Node 12' - v=0 ft/s - Elastic - Impact_Deck-Node 12' -v=3 ft/s - Elastic 

- Impact DeckNode 12' - v=3 ft/s - Plastic ^“Load Scaled 


Figure 3.12 Dynamic response of the rotational spring at node 12 and 12’. 


Flexible Wall 

Node 12 - 12' - Rotation - Z 

(Load applied at node 11 moving at the specified velocity "v") 
(kr)l = 12,426,909 : (kr)2 = 12,426,909 
(kr)l = 12,426,909 : (kr)2 = 6213.454.52 



Time (sec) 

- Impact_Beam-Node 12 - v=0 ft/s - Linear • • • • Impact Deck-Node 12' - v=0 ft/s - Linear 

-Impact Deck-Node 12' - v=3 ft s -Linear _ _Impact Deck-Node-12' - v=3 ft s - Nonlinear 































































ERDC/ITL TR-16-1 


69 


Figure 3.13 Dynamic response of the transverse spring located at x = 

84.5 ft. 


No n lin ea r Fo rce-D isp la cemen t 

^elastic = 0 003 ft : kl = 132,189.5 kip/ft: k2 = 66,094.75 kip/ft : k nnhMld = k 1 



0 0.002 0.004 0.006 0.008 0.01 

Lateral Displacement (ft) 


Nodal outputs provided from the FEO analysis of a flexible wall are the 
longitudinal displacement, transverse displacement, and rotational 
displacement (in radians) for each node at each time step of the simulation. 
A table is also provided that gives the maximum displacements 
(longitudinal, transverse, and rotational) for each node and the time that 
the maximum displacement occurs. 

Figure 3.14 shows the GUI table of maximums for the example problem in 
this section. From this GUI table, it is possible to tell the time step and the 
node with the maximum displacement for transverse, longitudinal, and 
rotational displacements (as shown in Table 3.4). 

Figures 3.15, 3.16, and 3.17 show the time histories for displacements at 
nodes 21 and 22 (longitudinal and rotational). Because some of these 
displacements are very small, and the data were stored with limited 
precision, some of these plots develop jaggies, where the data seems to 
form stair steps. 

The Impact_Deck GUI also allows the user to visualize the entire beam in 
motion using an exaggerated plot of displacements transversely, longitudi¬ 
nally, and rotationally. Figures 3.18,3.19, and 3.20 show the displacements 
of the wall from these animated plots at the moment where the maximum 
displacement occurred: 0.252 sec for transverse displacements, 0.192 sec 
for longitudinal displacements, and 0.22 sec for rotational displacements. 
These data were also subject to the jaggies because of the precision of the 
stored data. The data were scaled to fit the plot. 


































ERDC/ITL TR-16-1 


70 


Figure 3.14 lmpact_Deck GUI table of maximum nodal displacements for the 

McAlpine flexible wall. 



Table 3.4 Maximum nodal displacements for the McAlpine flexible wall example problem. 



Node number 

Value 

Time (seconds) 

transverse 

21 

0.0659 feet 

0.252 

longitudinal 

22 

0.0016 feet 

0.192 

rotational 

22 

0.0042 radians 

0.220 


Figure 3.15 Transverse nodal displacement time histories for node 21. 



























































ERDC/ITL TR-16-1 


71 


Figure 3.16 Longitudinal nodal displacement time histories for node 22. 


FEO Nodal Output 



** 

File 

FlexibleWallTestl .feo 

© Display extreme values and their times 




® Plot displacement vs. time Node Index 122 

▼ Displacement/Rotation: |y 



© Plot animated I 0 - 00000 

|4j| 0 Loop Displacement/Rotation: |x ▼ ! 




Y » 

u 


Node =22 Y Displacement 


-0.0007 

disp. (ftj 



Figure 3.17 Rotational nodal displacement time histories for node 22. 


FEO Nodal Output 

1 Rle 


" FlexibleWallTestl.feo 


1 © Display extreme values and their times 

| (®) Plot displacement vs. time Node Index 22 

▼ Displacement/Rotation: |z ▼ 

I © Plot animated Hj[MlfyifTT(HlWll I 0 - 00000 

|4j| 0 Loop UisplacementfRotation: |x »| 


Y • 

Ll: 



2.0 2.5 

time (s) 




















































ERDC/ITL TR-16-1 


72 






















































ERDC/ITL TR-16-1 


73 



Element outputs provided from the FEO analysis of an impact deck are the 
axial force, shear force, and moment for each element at each time step of 
the simulation. A table is also provided that gives the minimum and 
maximum forces and moments for each element and the times that the 
minimum and maximum force/moment occur. 

Figure 3.21 shows the GUI tables of extreme values for the example 
problem in this section. The beginning of the axial force extremes table is 
shown. The shear force extremes table and moment extremes table are 
available by scrolling in the interface. From this GUI table, it is possible to 
tell the time step and the element with the extreme axial force, shear force, 
and moment (as shown in Table 3.5). 

Because the simulated barge train impact occurs between nodes shared by 
elements 20 and 21, the extreme forces are found primarily in those 
elements. Figures 2.44 - 2.49 show the time histories for the axial force, 
shear force, and moments for the elements with the minimum and 
maximum values, respectively. 

Figures 3.28 - 3.30 show the axial force at 0.2 sec, shear force at 0.22 sec, 
and moment at 0.220 sec for the entire wall. These are created from the 
Impact_Deck GUI feature to view the animated forces. 































ERDC/ITL TR-16-1 


74 


Of primary importance for doing the pile group founded flexible wall 
analysis, is being able to see exactly how much force each pile group will be 
able to resist during an impact event. This is measured by finding the 
resisting force from the spring model used for the pile group for the pile 
group’s displacement. Figure 3.31 shows the table of forces/moments 
resisted at the nodes representing a pile group in the Impact-Deck 
example model. 

Figure 3.21 lmpact_Deck GUI table of element minimum and maximum axial 
forces for the McAlpine flexible wall example. 



Table 3.5 Extreme forces/moments for the lmpact_Deck example problem. 




Element number 

Value 

Time (seconds) 

Axial 

min 

21 

-332.80 kips 

0.200 

max 

20 

332.80 kips 

0.200 

Shear 

min 

21 

-541.18 kips 

0.220 

max 

20 

541.18 kips 

0.220 

Moment 

min 

31 

-1,036.58 kip-feet 

0.198 

max 

21 

3,111.78 kip-feet 

0.220 



































ERDC/ITL TR-16-1 


75 


Figure 3.22 Axial-force time histories for element 21. 





File 

FteodNeWalTest 1 ieo 



I O Display extreme values and their times 

(®) Plot force/moment vs. time Element Index [21 

▼ | Mode: [Axial 


Plot animated ^ || 3 

|^J| B Loop Mode: |Axial 


0 

Elem = 21 Axial Force 



Y » 

u 



-325.0 

- 332 - 79 «f.o 








































ERDC/ITL TR-16-1 


76 



Figure 3.25 Shear-force time histories for element 20. 


FEO Element Output 


S' 


HexibleWaBTest 1 ieo 


© Display extreme values and their times 

® Plot force/moment vs. time Element Index [20 Mode: [shear 

Plot animated M ► II N O-OOCOO 

a Elem = 20 Shear Force 

















































ERDC/ITL TR-16-1 


77 


Figure 3.26 Moment time histories for element 31. 


File 

RexiWeWalTest 1 ieo 




I O Display extreme values and their times 




(®) Plot force/moment vs. time Element Index [31 

▼ | Mode: |Moment 



Plot animated ^ || ^ M : 

^10 Loop Mode: |A»al 




Y » 

u 


force ( 



Figure 3.27 Moment time histories for element 21. 


FEO Element Output 


S' 


HexibleWaBTest 1 ieo 


© Display extreme values and their times 

® Plot force/moment vs. time Element Index [21 Mode: | Moment ▼ | 

O Plot animated [HHHEUKHIH] l 0 00000 

a Elem = 21 Moment 


















































ERDC/ITL TR-16-1 


78 


Figure 3.28 Wall axial forces at 0.2 sec. 


FEO Element Output 


ReodWeWaBTest 1 .feo 


© Display extreme values and their times 

© Plot fore e/moment vs. time Element Index; [ 31 Mode: [Moment 

® Plot animated fMHIfFIfJlIfMlfWI It 1 - 20000 &l Bl ^ Mode: (* 




3C C.332.8 


Figure 3.29 Wall shear forces at 0.220 sec. 


FEO Element Output 


FtedbleWaSTest 1 feo 


© Display extreme values and their times 
© Plot force/moment vs. time Element Index [31 


] Mode: [Moment ▼] 

~ g| [3 Loop Mode: [shear 




Shear Force (kips) 















































ERDC/ITL TR-16-1 


79 


Figure 3.30 Wall moments at 0.220 sec. 



Figure 3.31 Table of pile group response maximum displacements. 




















































ERDC/ITL TR-16-1 


80 


Figure 3.32 Response forces for the pile groups at time 0.2200 sec. 



Figure 3.33 Response forces for the pile groups at time 0.2800 sec. 






























































ERDC/ITL TR-16-1 


81 


Figure 3.34 Pile group response for the pile group at node 2 and at time 0.298 sec. 



Figures 3.32 and 3.33 show the instantaneous response forces at time 0.22 
and 0.28 sec, respectively. At time 0.22, the longitudinal and moment 
forces reach their peak values of 438.186 kips and -98.047 kip-ft for node 
44, which was created at the center of rotation for the pile group. At time 
0.28, the peak transverse response force is 638.156 kips at node 44. 

Figure 3.34 shows the response of the pile group at node 44 at time 
0.28 sec as a force-versus-displacement plot, and as force and 
displacement time histories. This is the time when the displacement for 
the pile group at node 44 is at its maximum location in the transverse 
direction. 

The Case for Dynamic Analysis: 

This example problem demonstrates the necessity of performing a 
dynamic analysis of these pile founded walls. This example uses an impact 
time history that is the result of the Winfield Test #10. This impact time 
history is applied transverse to the approach wall at a position starting at 
84.5 ft along this section of wall (close to the longitudinal location of node 
44) and moved at 1 ft/sec along the approach wall. The peak force for the 
impact time history was 516.4 kips at time 0.2 sec. 
























ERDC/ITL TR-16-1 


82 


Node 44 (as discussed above) has a maximum transverse force of 638.156 
kips, which occurs at 0.28 sec. This is the pile group node between two 
flexible walls, so its transverse displacement has been affected by the 
inertial effects of the walls and the over structure, as well as rotational 
response. The total pile group response for the section of flexible wall 
structure (with 3 pile group nodes) at this time (0.28 sec) is 662.269 kips, 
this total response force is greater than the peak input force. It is also the 
peak overall response. Table 3.6 shows the values generated by this data 
collection. Figure 3.35 takes this information a step further by displaying 
the time histories for the three forces. 


Table 3.6 Transverse forces with respect to time. 


Time (s) 

Node 44 Force (kips) 

Impact Load Force (kips) 

Total Pile Group 
Response (kips) 

0.200 

492.3062 

516.400 

483.616 

0.280 

638.1564 

446.358 

662.269 


Figure 3.35 Time-history plot of transverse input forces, total force response for all the pile 
groups, and an individual pile group response forces. 



Table 3.6 and Figure 3.35 show that the overall pile group response is 
dynamic because it does not track with the input force. Instead, the pile 
groups respond to the input force over time due to inertial effects. Because 
of this, the overall pile group response can have higher peak forces. For 























































ERDC/ITL TR-16-1 


83 


example, at the time of peak response of 0.2 sec a maximum impact force 
of 516.4 kips is applied to the approach wall. The total transverse force 
response of the 3 nodes representing the pile groups at the beam supports 
is 483.616 kips. This 6.3% smaller summed force response is due to the 
contribution of the first two terms of the equation of motion (also seen in 
Appendix A) for the dynamic structural response: 

[Af]{ii(f)} + [C]{li (t)} + [*]{u(t)} = {F(t)} (2.1 bis) 

The summed transverse effects are smaller than the peak load because the 
mass of the beams are accelerated slowly. At time 0.28 sec, when a peak 
response force is recorded at node 44 for an individual pile group, the 
contribution of the first two terms of the equation of motion for the 
dynamic structural response is even larger resulting in an overall response 
force of 662.269 kips. This is nearly 28.3% greater than the peak input 
force of 516.4 kips. These observations demonstrate the importance of 
applying the equation of motion for calculating pile group structural 
response forces (and displacements). These differences explain why a 
dynamic analysis is required versus a static analysis in which the user- 
provided impact load is applied as a single peak value (e.g., determined to 
be the input peak force from the time history). 

For simply supported beams, the effect of load sharing between pile 
groups is altered by the modal characteristics of the system (Ebeling et al. 
2012). Because this system is very stiff, the inertia of the beam and pile cap 
superstructure cause localized response at the node closest to the impact. 
This is shown by the fact that the peak response force at node 44 is 
638.156 kips, which is within 4% of the total transverse response force 
(662.269 kips). Notice that the peak transverse response force at node 44 
is greater than the peak transverse input force. This implies that dynamic 
analysis using impulse momentum principles should be performed to 
determine the greatest forces acting at any pile group. 

Validation Using Impulse Calculations: 

The higher peak values of the overall pile group response seem out of place 
until an impulse calculation (taking the area beneath the time history 
curves for input load and overall pile group response in Figure 3.35) is 
performed. Despite inertial effects, the impulse of the input load must be 
equivalent to the overall pile group response, if the piles do not fail. When 



ERDC/ITL TR-16-1 


84 


an impulse calculation is performed for the overall pile group response, 
the result is 463 kip-sec. The impulse for the input force is 463 kip-sec. 

The difference is minimal, less than 1 %. 

Load Sharing: 

For this type of flexible approach wall structural system, there can be load 
sharing (depending upon the structural detailing) in the longitudinal 
direction starting with the first pile group beyond the point of impact. 
There will also be load sharing in the transverse direction among the pair 
of pile bents supporting the impact beam for an impact anywhere along 
the simply supported beam. However, this structural configuration does 
not have the advantage of the significant load sharing among pile groups 
that the Lock and Dam 3 impact deck configuration possesses. This is 
exemplified by the observation that the node 44 maximum transverse 
force of 638.156 kips, which occurs at 0.28 sec, is greater than the peak 
input force of 516.4 kips occurring at 0.2 sec. The pile group total 
transverse response force is greater than the peak input force. For Lock 
and Dam 3, the peak transverse force for the pile group possessing the 
maximum peak force of any of the 96 pile groups was 56.4652 kips. The 
Lock and Dam 3 dynamic structural response analysis was subjected to the 
same input impact force time history specified in this analysis. 

3.9 Final Remarks 

In this section, the McAlpine flexible wall physical model was presented and 
the mathematical model to calculate the dynamic response was also 
developed. Impact_Deck is a computer program that was used to calculate 
the dynamic response of an elastic beam supported over linear elastic or 
plastic spring supports. The mathematical formulation modeled the ends of 
the simply supported beams with no moment transfer. The impact normal 
and parallel concentrated external load was located at a specified location or 
assumed to have motion at a specified constant velocity. The damping effect 
was considered by means of the Rayleigh damping model which depended 
on the natural frequencies of the system. These natural frequencies were 
calculated in an approximate way by using the linear stiffness of the pile 
groups and the mass of the impact beams. The results of Impact_Deck 
proved to be valid when compared to the results obtained with SAP2000. 
Finally, an example was presented to show the plastic behavior of the 
springs and how this result compared to the linear elastic response. 



ERDC/ITL TR-16-1 


85 


4 Traditional Impact Beam Guard Walls 

4.1 Introduction 

This chapter summarizes an engineering methodology using the 
Impact_Deck software for performing a dynamic structural response 
analysis of a flexible impact beam supported over pile groups and subjected 
to a barge impact event. For this example, a traditional model of impact 
beams simply supported on bents with in-line pile groups is examined. 

This example models a glancing blow impact event of a barge train 
impacting an approach wall as it aligns itself with a lock is an event of 
short duration; the contact time between the impact corner of the barge 
train and the approach wall can be as short as a second or as long as 
several seconds. The next generation of Corps approach walls is more 
flexible than the massive, stiff-to-rigid structures constructed in the past in 
order to reduce construction costs as well as to reduce damage to barges 
during glancing blow impacts with lock approach walls. A flexible 
approach wall or flexible approach wall system is one in which the wall has 
the capacity to absorb impact energy by deflecting or “flexing” during 
impact, thereby affecting the dynamic impact forces that develop during 
the impact event. 

4.2 Guard Walls - Physical Model 

Guard walls are a kind of flexible wall commonly used at locks by the 
Corps. Each segment of a flexible guard wall structure consists of a 
continuous elastic concrete beam with a span of approximately 50 or 60 ft 
long, each segment. The continuity of the beam is achieved by means of 
shear key at each pile group support. This means the beams transfer the 
longitudinal and transverse forces with no moment transfer at each pile 
supports. The axial and transverse forces at the end of the beams are 
transferred to the pile cap by means of a shear key. The shear key is a 
concrete block behind the end and start of two consecutive flexible impact 
beams. The length of the shear key is equal to the width of the pile cap of 
the pile group. The shear key is part of the massive pile cap that rest over 
the pile group. The pile group consists of two aligned piles, each with a 
diameter of 5 ft 8 in. The two piles are arranged in such a way that no 
torsion transfers to the pile group. A plan view drawing and a cross-section 
view are presented in Figure 4.1. 



ERDC/ITL TR-16-1 


86 


Figure 4.1 Guard wall schematic drawing. 



4.3 Guard wall - Mathematical Model 

The guard wall can be considered as a beam element because its length is 
much greater than the other two directions. The length of each segment is 
around 50 or 60 ft, heights and widths of 6 ft 7 in., and 9 ft, respectively, 
are not untypical. The mathematical model is described in Figure 4.2. The 
model in the analysis has two consecutive beams with longitudinal and 
transverse elastic-plastic spring supports at the start, mid span, and end of 
the system. These three sets of springs model the pile group at the start, 
mid span, and end of the two consecutive beams. The nodes that connect 
the beams to the pile cap do not transfer moment between the beams and 
the center pile cap. The center pile group is modeled with three springs 
(two translational and one rotational) in the generalized Impact_Deck 
software formulation. This model is similar to the McAlpine alternative 
flexible approach wall model but with two differences. First, no rigid link 
is used in the guard wall model. Second, no rotational elastic-plastic 
rotational spring is included. 

The mathematical model can be done using 3-D beam elements. A 3-D 
beam element has 6 degrees of freedom per node, producing 12 degrees of 
freedom per element. The degrees of freedom per node are 3 translations 
and 3 rotations as shown in Figure 4.3. The applied normal force F x (t) is 
the impact-force time history developed using the PC-based software 
Impact_Force. The applied parallel force F y (t) is a fraction of the normal 
force calculated using the dynamic coefficient of friction between the barge 
and the impact deck surfaces. 



















































ERDC/ITL TR-16-1 


87 


Figure 4.2 Guard flexible approach wall mathematical model. 



Figure 4.3 (a) Typical 3-D segment of the Impact Deck beam element, (b) Impact force applied 
to the Impact Deck, (c) Typical 3-D beam element. 



If the model used to describe the beam is developed in the plane, the beam 
element has 3 degrees of freedom per node and 6 degrees of freedom per 
element. The degrees of freedom per node are 2 translations and 1 
rotation, as shown in Figure 4.4. Based on the notation of Figure 4.3, the 
force and moment conditions for node i are Fi, x = Vi, Mi, x = o, Fi, y = Fi, Mi, y 
= o, Fi, z = o, and M;, z = Mi, and for node/ are Ff, x = Vf, Mf, x = o, Ff, y = Ff, Mf, y 
= o, Ff, z = o, and Mf z = Mf. Basically, to transform a 3-D beam element to a 
2-D (plane element), the moment about the “x” axis, the moment about 
the “y” axis, and the force in the “z” directions are equal to zero. 




































































ERDC/ITL TR-16-1 


88 


Figure 4.4 Typical 2-D beam element used in lmpact_Deck. 



The behavior of the three piles under static lateral load was done to 
determine the stiffness coefficient of the springs. To have an idea of the 
magnitude of the linear translational spring stiffness, refer to the 
calculation in Appendix E. 

4.4 Nonlinear force-deflection relationship for the springs supports 

Section 2.5 discussed how the nonlinear spring response force-displacement 
backbone curve was used to model plastic deformation in the pile 
substructure for dynamically loaded structures. Push-over calculations were 
similarly performed for the pile layout for guard wall supports. 

Figure 4.5 shows the results for a fixed-head single pile analysis from 
Figure 3.21 of Ebeling et al. (2012). Notice that these curves have three 
linear segments with two breakpoints. As the transverse loading at the pile 
cap increases, the bending moment at the top of pile-to-bent will increase 
until this moment connection yields and fixity is lost. After this occurs, the 
top of pile-to-bent behaves as a pinned-head condition with no constraint 
against rotation being offered within this region. Observe in the push-over 
curve that the rate of deformation has increased for the same incremental 
load after this hinge is formed. This results in a “softer” spring stiffness 
representation in this zone of the push-over curve. The pile bent system 
continues to resist the increase in lateral loading up until the level of loading 
that induces a second plastic hinge (shown as the second breakpoint in 
Figure 4.5). This second breakpoint is reached when the piles start to hinge 
at or below the mudline. Beyond this point, the push-over curve continues 
to provide the same resistance for a time until the plastic hinge rotation 
capacity of the piles are exhausted at this point below the mudline (refer to 
section A. 10 in Appendix A of Ebeling et al. 2012 for these this capacity 
computation) and can no longer support the structure. 

The resulting curve for the Saul analysis (CPGA) on the wet site for two 
DIP piles is shown in the force-versus-deflection, push-over curve values 
listed in Table 4.1. 













ERDC/ITL TR-16-1 


89 


Figure 4.5 Force-displacement relations from push-over analysis of a single guard wall pile. 



Displacement (inches) 


Table 4.1 Primary loading curve for the transverse spring model for a bent with two vertical piles. 


Force 

Deflection 

Notes 

(kips) 

(inches) 

(feet) 

Adapted from Appendix A of Ebeling et al. 
(2012) 

0.0 

0.0 

0.0 


810.0 

11.28 

0.94 

Pile to pile cap moment capacity reached 

980.0 

20.272 

1.68933 

Flexural plastic hinges develop in piles 
below mudline 

980.0 

26.272 

2.18933 

Plastic hinge rotation 


In the longitudinal direction, a push-over analysis must be performed for 
the pinned-head single pile condition, since there are no other piles to 
constrain the bent against rotation in the direction of the longitudinal 
load. The pile bent will maintain the same relative position with the top of 
the piles. The pile bent will rotate with the top of piles as the moments 
increase in the piles and will continue to rotate until the piles begin to 





























ERDC/ITL TR-16-1 


90 


hinge below the mudline. In this case, the push-over analysis was 
performed with COM624G as a single pile model because the piles were 
vertical and a pinned-head boundary condition was imposed at the top of 
the pile. Again, because two piles were used, the force acting due to 
deflection was doubled. 


Table 4.2 Primary loading curve for the longitudinal spring model for a bent with two vertical piles. 


Force 

Deflection 

Notes 

(kips) 

(inches) 

(feet) 

Adapted from Appendix A of Ebeling et al. 
(2012) 

0.0 

0.0 

0.0 


418.0 

31.1 

2.59167 

Flexural plastic hinges develop in piles 
below mudline 

418.0 

37.1 

3.09167 

Plastic hinge rotation 


4.5 Solving for the motion of the structure 

The equations of motion for a structure comprised of decks supported on 
groups of piles and their end-release computations similar to the guard 
wall model is given in Appendix A. Appendix I gives a discussion of 
Rayleigh damping with section 1 .3 giving information specific to the guard 
wall model. The numerical method to be used, either HHT-a or Wilson -0 
are discussed in Appendix F and G, respectively. 

4.6 Validation of lmpact_Deck Computer Program 

The validation of the Impact_Deck computer program for the guard wall 
model was made against the results obtained from the well-known 
computer program SAP2000. The beam had a total length of 100.0 ft. In 
that validation, the beam was modeled with 11 nodes and 10 beam elements. 
A set of linear elastic springs were located at node 1, 6, and 11 where the pile 
supports were placed. The strength of the concrete was assumed as/c = 
5000 psi producing a modulus of elasticity for the concrete of E = 
580393.25 ksf. The beam cross-sectional area and the beam second 
moment of area (moment of inertia) were 54.668 ft 2 and 517.2 ft 4 , 
respectively. The mass per linear foot of beam was calculated as m = 
0.25486 kip *sec 2 /ft. A damping factor of 0.02 or 2% of the critical damping 
was used in both computer programs. The force time history was the 
Winfield test # 10 (Ebeling et al. 2010), as shown in Figure 4.6 and applied 
at node 6 with zero translational velocity. The tangential-force time history 
was the same as the transverse but multiplied by a dynamic coefficient of 




ERDC/ITL TR-16-1 


91 


friction of 0.5. The translational spring stiffness was constant and equal to 
88,125 kip/ft. Figures 4.7 and 4.8 show the dynamic-response time histories 
for node 1 and node 6 in the transverse direction. Both computer programs 
presented basically the same response of the system. At node 1, some 
differences in magnitudes were apparent. However, the magnitudes were 
very small, which was associated to numerical approximations. 



Figure 4.7 Validation of lmpact_Deck against SAP2000 - Transverse 
displacement at node 1. 


Guard Wall 
Node 1 - X Direction 



- ImpactDeck-Linear — SAP2000-Linear 




















































ERDC/ITL TR-16-1 


92 


Figure 4.8 Validation of lmpact_Deck against SAP2000 - Transverse 
displacement at node 6. 


Guard Wall 
Node 6 - X Direction 



4.7 Numerical Example of the Elastic-Plastic Nonlinear Response 
Using lmpact_Deck 

In this section, a numerical example will be shown that demonstrates the 
“activation” of the plastic behavior of the pile group springs. This can be 
possible if the limit elastic displacement specified by the user was low 
enough to force the springs to enter into the plastic response zone. The 
input data for the Impact_Deck computer program for the guard wall were 
as follows. The beam has a total length of 100.0 ft long with 11 nodes and 
io beam elements. A set of linear elastic springs were located at node 1, 6, 
and n where the start, mid span, and end pile supports were placed. The 
strength of the concrete was assumed as fc = 5000 psi producing a 
modulus of elasticity for the concrete of E = 580393.25 ksf. The beam 
cross-sectional area and the beam second moment of area (moment of 
inertia) were 54.668 ft 2 and 517.2 ft 4 , respectively. The mass per linear foot 
of beam was calculated as m = 0.25486 kip *sec 2 / ft. A damping factor of 
0.02 or 2% of the critical damping was used in both computer programs. 
The force time history was the Winfield test # 10 as shown in Figure 4.6 
and applied at node 6 with zero translational velocity. The tangential-force 
time history was the same as the transverse but multiplied by a dynamic 
coefficient of friction of 0.5. The translational springs stiffness were equal 
to ki = 88,126.33 kip/ft and k 2 = 44,063.165 kip/ft with a stiffness for 
unload after the elastic displacement equal to kunioad = ki. The elastic 
displacement that defines the elastic and plastic zone was Seiastic = 0.003 ft. 
































ERDC/ITL TR-16-1 


93 


Figure 4.11 shows the dynamic-response time history obtained for node 12 
and 12’ (i.e., centerline of center pile group). It was observed that the 
permanent lateral displacement of approximately 0.0035 ft when the pile 
entered the plastic behavior. Figure 4.9 shows the transverse-displacement 
time history of node 1 for the elastic and plastic behavior. It was presented 
that no permanent displacement was reached by the transverse spring at 
node 1. For the two spring models, it remained in the elastic zone. 

Figure 4.10 shows the transverse-displacement time history of node 6 for 
the elastic and plastic behavior. A permanent displacement was reached by 
the transverse spring at node 6. For the elastic-plastic spring at node 6, 
permanent displacement of about 0.0035 ft was calculated. These 
behaviors can be observed in Figure 4.11, where the plastic response was 
reached (second slope in the force-displacement diagram) ending with a 
permanent displacement of around 0.0035 ft- It was important to observe 
the two stages where the spring load and unload in the plastic zone 
occured. That happened at an approximate time of 0.1 and 0.2 sec. 

4.8 lmpact_Deck GUI results 

The Impact_Deck GUI was also used to run the guard wall problem in this 
section. This section does not provide an engineering analysis, but gives an 
idea of what information was provided so that an engineering analysis 
might be made. 


Figure 4.9 Dynamic transverse response of node 1. 


Guard Wall 
Node 1 - X Direction 


0.001 

0.0008 

0.0006 

0.0004 
■g 0.0002 

4 0 

- 0.0002 

- 0.0004 

- 0.0006 

- 0.0008 

0 0.5 1 1.5 2 2.5 3 3.5 4 

Time (sec) 

— Impact_Deck-Elastic ImpactDeck-PIastic 


































ERDC/ITL TR-16-1 


94 


Figure 4.10 Dynamic transverse response of spring at node 6. 


Guard Wall 
Node 6 - X Direction 



ImpactDeck-Elastic Impact-Deck-Plastic 


Figure 4.11 Plastic force-displacement of the transverse spring at 

node 6. 



The model input for material properties was essentially the same as 
entered in section 4.7 with a few exceptions. The spring models 
(transverse and longitudinal), for each pile group was returned to the 
values specified in section 4.4 and shown in Figure 4.12. 

The geometry for the beams was more highly resolved, with 50 beam 
elements per wall section and 51 nodes per beam. The pile group nodes 
were node numbers 1, 51, and 101. In all other respects, the input models 
were similar. 
























































ERDC/ITL TR-16-1 


95 


Figure 4.12 ImpactJDeck GUI pile group longitudinal and transverse spring model 

backbone curves. 



Nodal outputs provided from the FEO analysis of a flexible wall were the 
longitudinal displacement, transverse displacement, and rotational 
displacement (in radians) for each node at each time step of the simulation. 
A table was also provided that gives the maximum displacements 
(longitudinal, transverse, and rotational) for each node and the time that 
the maximum displacement occured. 

Figure 4.13 shows the GUI table of maximums for the example problem in 
this section. From this GUI table, it is possible to tell the time step and the 
node with the maximum displacement for transverse, longitudinal, and 
rotational displacements (Table 4.3). 

Figures 4.14, 4.15, and 4.16 show the time histories for the displacements 
at nodes 51 and 50. Because some of these displacements were very small, 
and the data were stored with limited precision, some of these plots 
developed jaggies. 

The Impact_Deck GUI also allowed the user to visualize the entire beam in 
motion using an exaggerated plot of displacements longitudinally, trans¬ 
verse, and rotationally. Figures 4.17, 4.18, and 4.19 show the displacements 
of the wall from these animated plots at the moment where the maximum 





































ERDC/ITL TR-16-1 


96 


displacement occurred, 0.332 sec for transverse displacements, 0.41 sec for 
longitudinal displacements, and 0.286 sec for rotational displacements. 
These data were also subjected to the jaggies because of the precision of the 
stored data. The data were scaled to fit the plot. 


Figure 4.13 lmpact_Deck GUI table of maximum nodal displacements for the 

guard wall. 



Table 4.3 Maximum nodal displacements for the guard wall example problem. 



Node number 

Value 

Time (seconds) 

Transverse 

51 

0.7706 feet 

0.332 

Longitudinal 

51 

0.2409 feet 

0.41 

Rotational 

50 

0.0147 radians 

0.286 

























ERDC/ITL TR-16-1 


97 


Figure 4.14 Transverse nodal displacement time histories for node 51. 





























ERDC/ITL TR-16-1 


98 


Figure 4.16 Rotational nodal displacement time histories for node 50. 



Figure 4.17 Transverse wall displacements at 0.332 sec. 
































ERDC/ITL TR-16-1 


99 


Figure 4.18 Longitudinal wall displacements at 0.41 sec. 



Figure 4.19 Rotational wall displacements at 0.286 sec. 



File 

GuardWallTest1_pushover.feo 


O Display extreme values and their times 

O Plot displacement vs. time Node Index i ~ Displacement/Rotation: |z 

O' Plot animated ^ || H HI 0 28600 \^\ g] Loop UisplacemenVRotation: |z 



Time = 0.28600 Rotation 


100.0 

| 







70.0 








Y (ft) 50 0 








30.0 




10.0 

j 


Y 




L_ x 





-(J .0147 -0.01 -0.005 0.0 0.005 0.01 0.0 

147 

1 — w 

s - 

Uz (rad) 


Element outputs provided from the FEO analysis of an impact deck were 
the axial force, shear force, and moment for each element at each time step 
of the simulation. A table was also provided that gives the minimum and 












































ERDC/ITL TR-16-1 


100 


maximum forces and moments for each element and the times that the 
minimum and maximum force/moment occured 


Figure 4.20 show GUI tables of extreme values for the example problem in 
this section. The beginning of the axial force extremes table is shown. The 
shear force extremes table and moment extremes table are available by 
scrolling in the interface. From this GUI table, it is possible to tell the time 
step and the element with the extreme axial force, shear force, and moment 
(as shown in Table 4.4). Because the impact occurs at the midpoint of the 
guard wall and doesn’t move for this example, the forces are symmetric 
about the impact point (node 51). 


Figure 4.20 lmpact_Deck GUI table of element minimum and maximum axial 
forces for the guard wall Example. 


FEO Element Output 


' File 

Guard WallTest Ifeo 

(®> Display extreme values and their times 

O Plot force/moment vs. time Elementlndex i 

Plot animated \i ► || N — — 

e: [ Axial 

[71 Looo Mode: Axial 

zJ 



Axial 

Axial 

Axial 

Axial 


* 


Minimum 

Minimum 

Maximum 

Maximum 


1 

Elem 

Force 

Time 

Force 

Time 


H 

; id 

(kip) 

(sec) 

(kip) 

(sec) 


0 

1 

-138.55 

0.174 

138.55 

0.174 



2 

-137.99 

0.174 

137.99 

0.174 



3 

-137.43 

0.174 

137.43 

0.174 



4 

-136.87 

0.174 

136.87 

0.174 



5 

-136.31 

0.174 

136.31 

0.174 



6 

-135.75 

0.174 

135.75 

0.174 



7 

-135.18 

0.174 

135.18 

0.174 



8 

-134.62 

0.174 

134.62 

0.174 



9 

-134.05 

0.174 

134.05 

0.174 



10 

-133.49 

0.174 

133.49 

0.174 



11 

-132.92 

0.174 

132.92 

0.174 



12 

-132.36 

0.174 

132.36 

0.174 



13 

-131.79 

0.174 

131.79 

0.174 



14 

-131.22 

0.174 

131.22 

0.174 



15 

-130.65 

0.174 

130.65 

0.174 



16 

-130.09 

0.174 

130.09 

0.174 



17 

-129.52 

0.174 

129.52 

0.174 



18 

-128.95 

0.174 

128.95 

0.174 



19 

-128.38 

0.174 

128.38 

0.174 



20 

-127.81 

0.174 

127.81 

0.174 



21 

-127.23 

0.174 

127.23 

0.174 



22 

-126.66 

0.174 

126.66 

0.174 



23 

-126.09 

0.174 

126.09 

0.174 



24 

-125.52 

0.174 

125.52 

0.174 



25 

-124.95 

0.174 

124.95 

0.174 



26 

-124.37 

0.174 

124.37 

0.174 



27 

-123.8 

0.174 

123.8 

0.174 





" '■' T ' 

- """ '»'» 





Table 4.4 Extreme Forces/Moments for the Impact Deck Example Problem. 




Element number 

Value 

Time (seconds) 

Axial 

Min. 

1&100 

-153.99 kips 

0.410 

Max. 

1&100 

153.99 kips 

0.410 

Shear 

Min. 

51 

-214.86 kips 

0.714 

Max. 

50 

214.86 kips 

0.714 

Moment 

Min. 

27&75 

-2532.67 kip- 
feet 

0.716 

Max. 

26&74 

2532.67 kip-feet 

0.716 

















































ERDC/ITL TR-16-1 


101 


Figures 4.21 - 4.26 show the time histories for the axial force, shear force, 
and moments for the elements 50 and 51, respectively. This shows the 
symmetry of the solution. 


Figure 4.21 Axial force time histories for element 51. 


FEO Element Output 


I File 

13uandWallTest1_pushover.feo 


© Display extreme values and their times 

(§) Plot force/moment vs. time Element Index [ 51 ▼) Mode: | Anal 

© Reanimated fHflHTMIlHlll pa»»— 

Elem 



Mode: | Axial | 

= 51 Axial Force 



Figure 4.22 Axial force time histories for element 50. 


SuandWallTest 1 _pushover.feo 


© Display extreme values and their times 

<§> Plot force/moment vs. time Element Index [ 50 ▼) Mode: | Axial 

Plot animated ft M Ml M M --- 


Y 

L 


Axial Force 


force (kips) 

















































ERDC/ITL TR-16-1 


102 


Figure 4.23 Shear force time histories for element 51. 



Figure 4.24 Shear force time histories for element 50. 


FEO Element Output 
File 


3uard Wall Test 1 _pushover.feo 


© Display extreme values and their times 
® Plot force/moment vs. time Element Index 1 50 
Plot animated \i ► || M 0.0QC00 


1 Mode: Ishev 


Shear Force 









































ERDC/ITL TR-16-1 


103 


Figure 4.25 Moment time histories for element 51. 



Figure 4.26 Moment time histories for element 50. 



Figures 4.27 - 4.29 show the axial force, shear force, and moment at 
0.410, 0.714, and at 0.716 sec, respectively, for the entire wall. These are 
created to view the animated forces in the Impact_Beam GUI. 








































ERDC/ITL TR-16-1 


104 


Figure 4.27 Wall axial forces at 0.410 sec. 



Figure 4.28 Wall shear forces at 0.714 sec. 


FEO Element Output 


GuardWallTest 1 _pushoverfeo 


© Display extreme values and their times 

© Plot fbrce/moment vs. time Element Index: [l Mode: [Axial ▼ ] 

« Plotanimated [KHTFlMlMlItWl l°™» l»t BUw M ° d * H"-I 

































ERDC/ITL TR-16-1 


105 


Figure 4.29 Wall moments at 0.716 sec. 



Of primary importance for doing a pile-group founded flexible wall analysis 
is being able to see exactly how much force each pile group will be able to 
resist during an impact event. This is measured by finding the resisting 
force from the spring model used for the pile group for the pile group’s 
displacement. Figure 4.30 shows the table of forces/moments resisted at the 
nodes representing a pile group in the Impact_Deck example model. 

Figure 4.30 Table of pile group response maximum displacements. 





















































ERDC/ITL TR-16-1 


106 


Figure 4.31 Forces at the three pile group nodes at time 0.332 sec. 



Figure 4.31 displays the values for the forces and displacements for the pile 
group at node 51 at time 0.332 sec. Figure 4.32 shows the response of the 
pile group at node 51 at time 0.332 sec as a force-versus-displacement 
plot, and as force and displacement time histories. This is the time when 
the displacement for the pile group at node 51 is at its maximum location 
in the transverse direction. 

Figure 4.32 Transverse pile group response for the pile group at node 51 and at 

time 0.332 sec. 

^PM^Grou^Respon^ ^ 

\ File 

I GuardWallTest1_pushover.feo 

I © Display extreme values and their times 

I © Display Spring Forces at Time: [0 3320 ;^j 

O Plot Animated [Hf|[RQ[^][||][Ml[Hf| [0.33200 fej| [g] Loop Pile Group Node Index: [51 -| Degree of Freedom: [x ^1 

Time = 0.33200 Node: 50 X-Axis 
















































ERDC/ITL TR-16-1 


107 


The Case for Dynamic Analysis: 

This example problem demonstrates the necessity of performing a dynamic 
analysis of these pile founded walls. This example uses an impact time 
history that is the result of the Winfield Test #10 (Ebeling et al. 2010). This 
impact time history is applied transverse to the approach wall at a position 
starting at 50.0 ft along this section of wall (at the longitudinal location of 
node 51) and moved at o ft /sec along the approach wall. The peak force for 
the impact time history was 516.4 kips at time 0.200 sec. Table 4.5 provides 
a summary of transverse forces and the times at which they occur as well as 
the magnitude of the impact force occurring at this same point in time. Peak 
forces at the three times of interest are shown in bold in this table. 


Table 4.5 Transverse forces with respect to time. 


Time (s) 

Node 51 Force (kips) 

Impact Load Force (kips) 

Total Pile Group 
Response (kips) 

0.200 

398.9343 

516.400 

232.6187 

0.332 

663.9989 

320.2922 

816.9583 

0.392 

606.7369 

174.0544 

934.5545 


Node 51 (as discussed above) has a maximum transverse force of 663.999 
kips, which occurs at 0.332 sec. This is the pile group node between two 
flexible approach wall impact beams, so its transverse displacement has 
been affected by the inertial effects of the walls and the over structure, as 
well as rotational response. The total pile group response for the section of 
flexible approach wall structural model (with 3 pile group nodes) at this 
time (0.332 sec) is 816.958 kips. 

The peak overall force occurs at a later time for the guard wall bents (at 
0.392 sec) than in the input force time history (0.2 sec). This maybe due 
to the natural frequency of the system with less stiff supporting piles but 
more analysis (e.g., considering modal analysis of the piles as discussed in 
Appendix E of Ebeling et al. 2012) will be required. The peak overall force 
occurs at 0.392 sec and have a value of 934.555 kips, exceeding the peak 
input force of 516.4 kips by a good margin. 


Figure 4.33 takes this information a step further by displaying the time 
histories for the three forces. 




ERDC/ITL TR-16-1 


108 


Figure 4.33 Time-history plot of transverse input forces, total force response for all the pile 
groups, and an individual pile group response forces. 


Time Histories of Transverse Load and Response 
for Guard Wall Model 



Input Load 

Overall Pile Group Response 

Maximum Individual Pile Group 
Response (Node 51) 


From Table 4.5 and Figure 4.33, it can be seen that the overall pile group 
response is dynamic because it does not track with the input force. 

Instead, the pile groups respond to the input force over time due to inertial 
effects. Because of this, the overall pile group response can have higher 
peak forces. For example, at the time of peak response of 0.2 sec a 
maximum impact force of 516.4 kips is applied to the approach wall. The 
total transverse force response of the three nodes representing the pile 
groups at the beam supports is 232.619 kips. This 55% smaller summed 
force response is due to the contribution of the first two terms of the 
equation of motion (Equation 2.1, also seen in Appendix A) for the 
dynamic structural response: 

M{ fi ( f )Wc]H f )} + MM0M F (0} (2-1 bis) 


The summed transverse effects are thought to be smaller than the peak 
input force because the mass of the beams is accelerated slowly. At time 
0.332 sec, when a peak response force is recorded at node 51 for an 
individual pile group, the contribution of the first two terms of the equation 












































ERDC/ITL TR-16-1 


109 


of motion for the dynamic structural response is even larger, resulting in an 
overall response force of 663.999 kips. This is nearly 28.6% greater than the 
peak input force of 516.4 kips (occurring at 0.2 sec). Additionally, the 
overall response does not reach a peak until 0.392 sec with an even larger 
value of 934.555 kips. This greater overall force results from the pile groups 
all developing a positive deformation at the same time. These observations 
demonstrate the importance of applying the equation of motion for 
calculating pile group structural response forces (and displacements). These 
differences explain why a dynamic analysis is required versus a static 
analysis in which the user-provided impact load is applied as a single peak 
value (e.g., determined to be the input peak force from the time history). 

For this type of simply supported impact beam structural system with two 
beams, the modal contribution characteristics of the substructure system 
(Ebeling et al. 2012) may be important. Because this pile substructure 
system is relatively flexible, the inertia of the impact beams and pile cap 
superstructure may cause vibrations to be more in sync with the impact 
event. Expansion of the current dynamic Impact_Deck model would be 
required to account for this feature. 

Validation Using Impulse Calculations: 

The higher peak values of the overall pile group response seem out of place 
until an impulse calculation (taking the area beneath the time history 
curves for input load and overall pile group response in Figure 3.35) is 
performed. Despite inertial effects, the impulse of the input load must be 
equivalent to the overall pile group response, if the piles do not fail. When 
an impulse calculation is performed for the overall pile group response, 
the result is 461 kip-sec. The impulse for the input force is 463 kip-sec. 

The difference is minimal, less than 1 %. 

Load Sharing: 

For this type of flexible approach wall structural system, there can be load 
sharing (depending upon the structural detailing) in the longitudinal 
direction starting with the first pile group beyond the point of impact. 
There will also be load sharing in the transverse direction among the pair 
of pile bents supporting the impact beam for an impact anywhere along 
the simply supported beam. However, this structural configuration does 
not have the advantage of the significant load sharing among pile groups 



ERDC/ITL TR-16-1 


110 


that the Lock and Dam 3 impact deck configuration possesses. This is 
exemplified by the observation that the node 51 maximum transverse force 
of 663.999 kips, which occurs at 0.392 sec, is greater than the peak input 
force of 516.4 kips occurring at 0.2 sec. The pile group total transverse 
response force is greater than the peak input force. For Lock and Dam 3, 
the peak transverse force for the pile group possessing the maximum peak 
force of any of the 96 pile groups, was 56.4652 kips. The Lock and Dam 3 
dynamic structural response analysis was subjected to the same input 
impact-force time history specified in this analysis. 

4.9 Final Remarks 

In this section, the flexible guard wall physical model was presented and 
the mathematical model to calculate the dynamic response was also 
developed. Impact_Deck calculated the dynamic response of an elastic 
beam supported over linear elastic or plastic spring supports. The 
mathematical formulation also modeled the center pile group connection 
to the ends of the simply supported impact beams with zero moment 
transfer. The impact normal and parallel concentrated external load can 
be located at a specified location or can be assumed to have motion at a 
specified constant velocity. The damping effect was considered by means 
of the Rayleigh damping model which depended on the natural 
frequencies of the system. These natural frequencies were calculated in an 
approximate way by using the linear stiffness of the pile groups and the 
mass of the impact beams. The results of Impact_Deck proved to be valid 
when compared to the results obtained with SAP2000. Finally, an example 
was presented to show the plastic behavior of the springs and how this 
result compared to the linear elastic response. 



ERDC/ITL TR-16-1 


111 


5 lmpact_Deck Graphical User Interface 
(GUI) 

5.1 Introduction 

This section introduces the Impact_Deck GUI, which provides pre¬ 
processing and post-processing capabilities to the Impact_Deck engineering 
code. The purpose of the GUI is to provide the user a way to specify the 
input model types, input model parameters, analyze the input, and visualize 
the output. 

The program has a simple menu that allows the user to create a new data 
input set, open an existing set of input data, save the current input data in 
an existing or new input file, and exit the program. On the line below the 
menu, enter the title of the project; this will provide a reference for the 
user. 

Beneath the title bar, the Impact_Deck GUI uses a tabbed data input 
scheme where input data were grouped by related data and functionality. 
The Introduction Tab shows cross-section and plan views of examples of 
the different types of structures that can be analyzed with the 
Impact_Deck software (Figure 5.1). 

Figure 5.1 Introducing lmpact_Deck. 












































ERDC/ITL TR-16-1 


112 


The following sections discuss the other tabs in the Impact_Deck program. 

5.2 Geometry Tab 

The Geometry Tab is made up of subsections of data that specify the 
positions and velocities associated with the wall model and the source of 
the impact (typically, a barge train). These positions and velocities are 
entered in feet and feet/second, respectively. The rest of the program 
assumes English units for version l.o. The geometry information assumes 
a right-handed coordinate system, with the lock approach wall lying along 
the Y-axis (i.e., the longitudinal direction) and the X-axis proceeds into the 
wall (i.e., the transverse direction). 

The first section, at the upper left corner of the tab, is the selection of the 
type of wall to analyze. The choices reflect the three types of walls 
discussed previously; flexible approach walls, guard walls, and impact 
decks. Selecting any of these options changes the inputs available for the 
rest of the program. Figures 5.2 through 5.4 show the Geometry Tab when 
each of the options is selected. 

Figure 5.2 Geometry for a flexible wall. 










































































ERDC/ITL TR-16-1 


113 


Figure 5.3 Geometry for a guard wall. 



Figure 5.4 Geometry for an lmpact_Deck. 



The second subsection is unaffected by the approach wall type. This is the 
load information subsection. In this subsection are the inputs for the Y- 
position for the start of the impact and the velocity at which the impact 































































































































ERDC/ITL TR-16-1 


114 


travels along the wall. The impact time history, which is input on the next 
tab, determines the duration of the analysis. 

For flexible approach walls and guard walls, only two simply supported 
impact beams on three pile groups are included in the model. The 
Additional Pile Groups section allows the user to specify the number of 
pile groups that exist prior to and after the modeled section, so that those 
pile groups can contribute to the response. 

An impact deck consists of a series of impact monoliths with each 
monolith supported by clustered groups of piles. The impact deck 
structure is modeled as an entire set of beams spanning between multiple 
pile supports. Each pile support is modeled as a pair of transverse and 
longitudinal nonlinear springs (established through a push-over analysis 
as outlined in Ebeling et al. 2012). The fixity for the Starting End 
subsection allows the user to specify how the starting end of the approach 
wall, at the end away from the lock, is affixed. The fixed support constrains 
the end point in translation and rotation. The simple support has the end 
of the wall section resting on the connection to the end cell. 

At the right of the tab is the nodal input section for the wall. Because the 
wall always proceeds along the Y-axis, all of the nodes specified by the user 
can be entered with only a Y-coordinate. This approach wall has its origin 
at the point along the approach wall that is furthest from the lock 
chamber. 

For the flexible approach wall and guard wall models, the first and last 
nodes in the wall are automatically assigned to be pile group nodes, but 
the user must flag one of the internal nodes as associated with the central 
pile group. The pile group nodes are the nodes where the spring models for 
the piles resist the impact on the wall. The process for creating the central 
pile group will be discussed further in this section. 

For the Impact_Deck model, most of the nodes are connected to a pile 
group, but some nodes represent only the connectivity between the 
sections of impact deck monoliths. These nodes are called inter-monolith 
nodes. The inter-monolith nodes are not connected to a pile group and 
have different end-release properties. The process for creating the inter¬ 
monolith nodes is similar to the method used to create the central pile 
group for the flexible wall and guard wall models. 



ERDC/ITL TR-16-1 


115 


Before discussing the creation and deletion of nodes, the visualization of 
nodes must be discussed. In the node input section to the right of the tab, 
there is a list containing the node information. The list will provide the 
position of the node and if the node is a central pile group/inter-monolith 
node or not. This list is given to provide the user with specific nodal data. 

At the bottom of the tab is the Input Plot area. This plot shows the existing 
nodes, their connectivity, load conditions on the wall, and the number of 
pile groups prior and after the displayed wall segment (for flexible wall 
and guard wall models). The position of the starting point of the barge 
train impact is shown with an arrow pointing at the wall. The direction of 
the barge train velocity (parallel to the approach wall) is input prior to 
specifying an impact-force time history. After an impact-force time history 
has been selected, the time history is displayed from the starting point 
until the end of the time history due to the velocity of the impact. 

Nodes are displayed with different colors depending on whether the pile is a 
pile group/inter-monolith node. Blue nodes represent unsupported nodes 
(no pile group) for the flexible approach wall and guard wall models and 
regular pile group nodes for the Impact_Deck model. The red nodes 
represent the pile group nodes for the flexible approach wall and guard wall 
models and inter-monolith connection nodes for the Impact_Deck model. 

When the mouse is moved across the Input Plot window with no mouse 
button pressed, the node that the cursor is closest to will be highlighted, 
and information about that node will be presented to the right of the Input 
Plot window. This is shown in Figure 5.3 and Figure 5.4. 

The view in the Input Plot window can be zoomed by click-dragging with 
the right mouse button (Figure 5.5). The view will be changed to display 
everything in the selected region with the aspect ratio maintained 
(Figure 5.6). The button to the right of the Input Plot window with a global 
map on it is the Zoom Extents button. Clicking this button reveals the entire 
wall as it is currently defined. 

Nodes can be selected by click-dragging with the left mouse button. Nodes 
that are surrounded by the dragged bounding box will all be selected. 
Multiple selection regions are not permitted at this time. Selected nodes are 
drawn with a line through them to differentiate them from the unselected 
nodes (Figure 5.7). Selected nodes maybe copied or deleted, as discussed 
below. 



ERDC/ITL TR-16-1 


116 


Figure 5.5 Zooming in the input plot section. 


Input Plot 




o Pile Group Node 

o Beam Node 

m 

Node 68 

r Y 

X 



YPos: 

415.9791666 


Figure 5.6 The zoomed view. 

Input Plot 

1 

3 

Node 86 

YPos: 

524.1666665 

r Y 

X 


Figure 5.7 Selected nodes are highlighted. 


Input Rot 


o i ^ 

■ 14 o 

Node 86 

- -- --- ° PTTtTttTtTTi 

r Y 

X 

' • T 0 

YPos: 

524.1666665 


There are three ways to create nodes in the node list. These three methods 
allow the user to input individual nodes, multiple nodes using 
interpolation, and copying and pasting nodes. 

The Single Node Input subsection allows the user to specify a location 
along the wall as a Y-axis position. Recall that the wall lies along the Y-axis 
and that the X-axis is into the wall with a right-handed coordinate system. 
A checkbox permits the user to specify whether this node is a central pile 
group node for the flexible wall and guard wall models and an inter¬ 
monolith node for the Impact_Deck model. Clicking the Add Node button 
in this subsection adds the node to the node list and plots it in the Input 
Plot area. Adding a node at the location where a node already exists will 
not create a new node, but can change the status of the node to or from a 
central pile group/inter-monolith node. 

Multiple nodes can be input using the Interpolated Node Input subsection. 
When a start and end position are entered with a number of divisions 
between nodes in that distance, nodes will be linearly distributed in that 
distance. The number of nodes placed will be equal to the number of 
divisions plus one; a node is placed at the start position and then the 
following nodes are placed at the total length divided by the number 
divisions away from the previous node, until the end point is reached. 

















































ERDC/ITL TR-16-1 


117 


Because there can be only one central pile group, interpolated nodes are 
not allowed to set that status. However, each node could be an inter¬ 
monolith node, so setting that status is allowed. Again, nodes that will be 
placed at the same location as existing nodes will not create a new node, 
but can change the status of the existing node. 

If there are nodes selected, then nodes can be created by clicking the Copy 
Selected Nodes button. When the button is selected, the Copy Selected 
Nodes dialog will appear that asks for an offset for the selected nodes 
(Figure 5.8). Clicking the Cancel button will terminate the copy event, but 
clicking Accept will cause the Copy Selected Nodes with Offset dialog box 
to open (Figure 5.9). Clicking the Accept button in this dialog creates a 
copy of the selected nodes with their attributes at the offset location 
relative to the original nodal positions (Figure 5.10). The original nodes 
are then deselected so they will not be copied again, because the Copy 
Selected Nodes with an Offset dialog box stays open in case multiple 
copies of the nodes needs to be made at the same relative distance. 

Clicking the Cancel button does not copy the last set of selected nodes, and 
terminates the operation. 

There are two methods for removing nodes. The Empty Node List button 
removes every node in the model. The Delete Selected Nodes button 
removes only the selected nodes, as shown in the Input Plot window. 

Figure 5.8 Entering an offset to copy 
selected nodes. 



Figure 5.9 Confirming the offset 
copy (which can be performed 
multiple times). 


Copy Selected Nodes with Offset 


Copy selected nodes with offset of 1000? 


OK || I Cancel 



































ERDC/ITL TR-16-1 


118 


Figure 5.10 Selected nodes are copied at the offset position. 



5.3 Impact Time History Tab 

The Impact Time History Tab is very simple. There is a method to browse 
for a time-history file and there is a button to adjust the time history to 
allow time for the pile-founded impact deck or flexible approach wall to 
reach a state where little-to-no displacements occur. 

The Browse button permits the user to bring in an impact time history file. 
Currently, the only format supported is the “.ETH” format output by 
Impact_Force. When an impact time history has been chosen the path for 
the file selected, relevant comments about the time history and actual 
values are displayed per Figure 5.11. 

The time history can be altered to add samples of zero force at the end to 
allow time for the dynamic structural response to settle after the 
deformations have completed. Clicking the Extend Force Time History 
button brings up the Extend Force Time History dialog. The user can 
specify a new length for the time history and the time history will be 
extended to the new length (Figure 5.12). 




























































ERDC/ITL TR-16-1 


119 


Figure 5.11 Input for an impact time history. 



Figure 5.12 Extending a time history with 
0.0 value. 



5.4 Beam Properties Tab 

The beam properties are different depending on whether a flexible 
approach wall model is chosen or if the guard wall or Impact_Deck model 
is chosen. For the flexible approach wall model, the beam properties also 
include the ability of the beam to flex and rotate differently than the guard 
wall beam or impact deck’s deck (Figures 5.13 and 5.14). 

The primary beam input, which is shared between the three models, are 
the Modulus of Elasticity, Cross-sectional Area, Moment of Inertia, Mass 
per Unit Length, Damping Ratio, and the Coefficient of Friction Between 
the Barge and the Beam/Deck. 











































ERDC/ITL TR-16-1 


120 


Figure 5.13 Beam properties tab as it appears for a flexible wall. 


a? Impact_Deck FlexibleWallTestl.idf 


File 

™e: FLEXIBLE WALL TEST 

Introduction Geometry Impact Time History [ Beam Properties] Pile Ouster Springs Analyze Output 


Beam Properties 


Modulus of Elasticity: 580393.25 
Cross-sectional Area: 54.6675 


Moment of Inertia: 517.1967 


Mass Per Unit Length: 0.25466 


Damping Ratio: 0.1 


Coefficient of Friction Between Barge and Deck: 0.5 


Beam Properties (Longitudinal) 

Modulus of Elasticity 
Cross-sectional Area 
Moment of Inertia 
Mass Per Unit Length 


Beam Properties (Transverse) 


Modulus of Elasticity 
Cross-sectional Area 
Moment of Inertia 
Mass Per Unit Length 


Length: 7.708 


kips/fT2 

fT2 

fT4 

(kips*sec"2)/ft 


kips/fT2 

fT2 

fT4 

(kips*sec~2)/ft 


kips/fT2 

fT2 

fT4 

(kips*sec"2)/ft 


Figure 5.14 Beam properties tab as it appears for an impact deck or guard wall. 


Tit'e: IMPACT DECK TEST PROBLEM 1 

Introduction Geometry Impact Time History [ Beam Properties! Pile Ouster Springs Analyze Output 


Beam Properties 


Modulus of Elasticity 
Cross-sectional Area 
Moment of Inertia 
Mass Per Unit Length 
Damping Ratio 


kips/fT2 

fT2 

fT4 

(kips*sec~2)/ft 


Coefficient of Friction Between Barge and Deck: 0.5 


Beam Properties (Longitudinal) 


Modulus of Elasticity: 
Cross-sectional Area: 

Moment of Inertia: 
Mass Per Unit Length: 



Beam Properties (Transverse) 


Modulus of Elasticity: 
Cross-sectional Area: 

Moment of Inertia: 
Mass Per Unit Length: 



Length: [(T 


kips/fT2 

ft~2 

fT4 

(kips*sec~2)/ft 


kips/fT2 

fT2 

fT4 

(kips’sec^yft 

ft 





































































ERDC/ITL TR-16-1 


121 


Flexible walls have input for properties acting along the beam and 
transverse to the beam. The longitudinal properties are the Modulus of 
Elasticity, Cross-sectional Area, Moment of Inertia, and Mass per Unit 
Length for the length of the wall. The transverse properties are the 
Modulus of Elasticity, Cross-sectional Area, Moment of Inertia, Mass per 
Unit Length, and the Length of the wall beam. 

5.5 Pile Cluster Spring Tab 

For the Impact_Deck software, pile clusters are modeled as non-linear 
springs. In the Pile Cluster Spring Tab, these non-linear pile cluster 
springs are defined. 

Again, there is a difference based on the wall model being used. Because 
the flexible approach wall model has pile groups that can rotate, entry of a 
rotational spring model is allowed (Figure 5.15). Figure 3.1 shows an 
example of a clustered three-pile group providing rotational resistance, 
with Appendix E summarizing the computations leading to the value 
assigned to its rotational spring stiffness. Because the guard wall and the 
impact deck wall models do not allow the pile caps to rotate, rotational 
springs are not defined (Figure 5.16). 

Figure 5.15 Beam properties tab as it appears for a flexible wall. 



































ERDC/ITL TR-16-1 


122 


Figure 5.16 Beam properties tab as it appears for an lmpact_Deck. 



To the right of this tab, there is a graph of the currently defined springs 
(longitudinal, transverse, and possibly rotational). Radio buttons next to 
these graphs allow the user to choose the spring definition currently being 
edited. The currently edited spring definition has a graph with a white 
background, and its values are displayed in the Edit subsection to the left. 

In the Edit subsection, there are three changes the user can apply to the 
current spring 

• Change the Unload/Reload characteristics of the curve 

• Add a force-displacement point to the curve 

• Delete a force displacement point from the curve 

There are two values that can be entered to adjust the Unload/Reload 
characteristics of the spring model. The first value is the amount of elastic 
lateral displacement allowed before plastic deformation occurs. The 
second value is the unload/reload stiffness, which is the force-versus- 
displacement slope that will be followed from the peak displacement and 
its force after a plastic deformation has occurred. These values will 
typically be the point where the first point on the curve is and the slope the 



































ERDC/ITL TR-16-1 


123 


first linear segment makes on the curve, but entry of these values lends 
more flexibility. 

Adding a point to the curve requires that the user enter a displacement and 
the force at that displacement, and then press the Add Point to List button. 
The points in the list (which are shown below the Point Entry subsection) 
are sorted by displacement. The points in the list are plotted in the current 
graph. Multiple forces may not be entered at the same displacement. When 
a new displacement/force pair is at the same displacement as an existing 
point, the new point overrides the previous point. 

To delete a point in the current spring model, the user can click on the row 
in the list for the point. That point is highlighted in blue. Pressing the 
Delete Point from List button removes the point from the list. 

These operations work for each of the spring model types. 

5.6 Analyze Tab 

The Analyze Tab (Figure 5.17) allows the user to: 

• View his wall model, 

• Choose a solution method, 

• Set the capture rate for data, 

• Select specific elements for more direct Finite Element Data, and 

• Perform an analysis. 

The input plot display mirrors the input plot display on the Geometry Tab, 
but does not allow the user to select nodes. In this way, a wall model can 
be quickly verified. 

There are two solution methods that can be applied to compute the 
dynamic structural response for the approach wall model: the HHT-a 
method and the Wilson -0 method. Each method has a different tolerance 
for the solution to be met. For the HHT-a method an alpha tolerance can 
be specified that must be met to finish the simulation. Similarly, for the 
Wilson -0 method, a value for Theta can be specified such that the method 
will converge. 



ERDC/ITL TR-16-1 


124 


Figure 5.17 Analyze tab input for analysis method and specified output. 



Time-history response analysis of a MDOF structural system model can 
result in large files of computed time histories for the various output 
variables. In order to reduce the size of the output files, the following feature 
has been added: each time step in the computed simulation results can be 
captured or the user can skip past a designated number of time steps. 
Entering l in the Capture every Time Step subsection guarantees that every 
time step is captured; entering 2 in the subsection means that every other 
time step is captured; and entering 3 in the subsection means that every 
third time step is captured and is repeated thusly. 

If the checkbox labeled Output Finite Element Data is checked, then the 
user can select specific finite element nodes that he/she would like to have 
text output for. The list next to the checkbox shows the nodes that will 
currently be output. To change the list, press the Manage Output button. 

Pressing the Manage Output button brings up a dialog box that gives the 
user the power to select output nodes (Figure 5.18). A diagram similar to 
the Input Plot of the Geometry tab is on the left of the Manage Output 
dialog. This diagram can be zoomed in much the same manner and a 
Zoom Extents button is provided. The node selection process works in a 








































ERDC/ITL TR-16-1 


125 


different fashion in the Manage Output dialog. When the user left click- 
drags an area in the diagram, the nodes are toggled between being selected 
and not selected. In other words, if a node was selected when the mouse 
was left click-dragged over it, it will be deselected. This allows the user to 
select multiple groups of nodes by selecting a group and then deselecting a 
subgroup. Selected nodes are shown highlighted. 

Figure 5.18 Selecting finite element nodes 
where data will be captured. 



A current list of selected nodes is displayed at the right of the dialog. There 
are also options to delete a node by selecting it from the list and then 
pressing the Remove Node button, and to clear the entire list by pressing 
the Clear List button. In order to use the list chosen in the Manage Output 
dialog, press the Accept button. Pressing the Cancel button keeps the 
original list. 

At the bottom of the Analyze Tab is the button labeled Perform Analysis. 
When this button is pressed, the processor is started with the current 
input data. 
































ERDC/ITL TR-16-1 


126 


5.7 Output Tab 

In the Output tab, output data files can be selected by the users for 
visualization of results. There are two main files that are output for each 
Impact_Deck processor run. There is the .FEO (Finite Element Output) file 
which contains the node and element results in a time domain solution, and 
there is the Run File output in .OUT format that explains how the run 
proceeded and provides only the requested node input from the Analysis 
Tab, so the user does not have to look for specific information in a very large 
file. 

The Output tab is therefore broken into two sections (Figure 5.19). Each 
section has a Load Output button, which is used to browse for the specific 
file. These files are automatically populated with the output from the 
current input file when the Perform Analysis button is pressed on the 
Analyze tab. When a file is selected for either section, that section’s 
buttons become enabled, and that file’s data is available to be visualized 
(Figure 5.20). 

Figure 5.19 Output tab for selecting and viewing select data. 






























ERDC/ITL TR-16-1 


127 


Figure 5.20 Output tab with selected data. 


Title: IMPACT DECK TEST PROBLEM 1 

Introduction Geometry Impact Time History Beam Properties Pile Ouster Springs Analyze | Output 


FEO Output Options 


Data Path : | lmpact_Deck_TestProbleml .feo 


Load Output- 


Show Nodal Information 


Show Element Information 


Show Pile Group Response 


Run Output Options 


Data Path : lmpact_Deck_TestProbleml .out 


Show Run Output 


For the FEO output file, there are three data to visualize: the Nodal 
Information, the Element Information, and the Pile Group Response. For 
the Run file output, the user can see the text file detailing how the program 
ran, and the specific finite element information selected on the Analyze 
Tab. These options are discussed below. 

5.7.1 FEO Nodal Output 

Pressing the Show Nodal Information button from the Output Tab brings 
up the FEO Nodal Output window. The window has a minimal menu, with 
only options to save the currently displayed information (either as text or 
as a .PNG bitmap graphic file for graphical data) and to close the window. 
The currently selected file name is displayed immediately below the menu. 
Beneath the menu are radio buttons that allow the user to present 
different information from the .FEO file. 

The FEO Nodal Output window presents nodal information in three ways 

1. The maximum values of all the nodes are presented as text 

2 . Individual nodal properties are shown as a 2-D plot of displacement versus 
time 





















ERDC/ITL TR-16-1 


128 


3. The entire structure is shown as a 2-D animated plot with the structure 
along the Y-axis, and the displacement along the X-axis. 

When the “Display extreme values and their times” radio button is selected, 
a list of nodes with each node’s maximum values of displacement (longi¬ 
tudinal, transverse, and rotation) and the times when those maximums 
occurred is displayed (Figure 5.21). The longitudinal displacements occur 
along the wall beam, transverse displacements occur into the wall beam, 
and rotations occur about the node in the beam. 


Figure 5.21 FEO nodal output window showing maximum nodal values for all the 

nodes. 



When the “Plot displacement vs. time” radio button is selected, the user 
can select a node to view using the “Node Index” combo box. When a node 
has been selected, the user can select whether to plot the node’s X- 
displacement, Y-displacement, or Rotation versus Time in the plot below 
by selecting from the “Displacement/Rotation” combo box. 

When these options have been selected, the graphic window below shows 
the relevant plot (Figures 5.22 through 5.24). At the top of the display is the 
selected node number and selected displacement. To the left of the plot is a 
display of the coordinate system and the wall with nodes and connectivity. 
The selected node is displayed in red to highlight where the node is in 






















ERDC/ITL TR-16-1 


129 


Figure 5.22 Graph of node 70 X-displacement vs time. 


FEO Nodal Output 


File 

lmpact_Deck_T estProbleml .feo 


© Display extreme values and their times 
# Plot displacement vs. time Node Index [70 

® P '°* a " imated jMTMlM P°000~ 


| Displacement/Rotation: [x~ 


Displacement/Rotation: [x ▼] 


Node =70 X Displacement 



Figure 5.23 Graph of node 70 Y-displacement vs time. 


FEO Nodal Output 


lmpact_Deck_TestProbleml .feo 


© Display extreme values and their times 
# Plot displacement vs. time Node Index [ 70 
: Plotan,mated ft M ► II M W 


Displacement/Rotation: 


[ 7 ] i_Qop Displacement/Rotation: [x =] 


Node = 70 


Y Displacement 


disp. (ftV 























































ERDC/ITL TR-16-1 


130 


Figure 5.24 Graph of node 70 Z-displacement vs time. 


FEO Nodal Output 


File 

lmpact_Deck_T estProbleml .feo 


© Display extreme values and their times 
# Plot displacement vs. time Node Index [70 

® fllSOTUMliMlWI P°°°0~ 


] Displacement/Rotation: (z 


Displacement/Rotation: |x ▼] 


Node = 70 Rotation 



relation to the wall. The rest of the area is the 2-D plot of the displacement 
versus time for the selected node and displacement. When the mouse is 
placed over this plot, a tooltip is shown revealing the displacement for the 
time that the cursor is over. 

It should be noticed that some displacements have some artifacts in their 
plots, typically shown as “staircase steps”. The “staircase steps” occur when 
the resolution of the data, with small numbers, outstrips the precision of the 
floating point variables used to represent the data. Aggregating this data 
leads to discontinuities in the curve which appear as “steps”. These steps 
occur in the low amplitude region of the curve which is inconsequential to 
the design. Notice in Figure 5.24 that the image depicts the vertical 
displacement of a node of the impact model and that the vertical displace¬ 
ment only varies by 0.0009 inches at the extremes, which will have an 
inconsequential bearing on the response forces of the structure. 

When the “Plot animated” radio button is selected, the user can select 
whether to plot the wall’s X-displacement, Y-displacement, or Rotation 
versus Time in the plot below by selecting from the “Displacement/ 


























ERDC/ITL TR-16-1 


131 


Rotation” combo box. The displacements/rotations at each node are linearly 
interpolated from node to node. 

Immediately next to the radio button is the animation control. This control 
has typical buttons that 

• Return the animation to the beginning time step, 

• Step to the previous time step, 

• Play the animation, 

• Pause the animation, 

• Step to the next time step, and 

• Take the animation to the last time step. 

Next to the buttons in the animation control is the current time display. 
The time display can be stepped up or down using the arrows next to the 
display. A checkbox next to the current time display allows the user to 
select if the animation will loop to the beginning or stop when the last time 
step has been reached. 

When these options have been selected, the graphic window below shows 
the relevant plot (Figures 5.25 through 5.27). At the top of the display is 
the current time and selected displacement. To the left of the plot is a 
display of the coordinate system and the wall with nodes and connectivity. 
The plot of the impact-load time history from the starting position of the 
impact to the ending impact location, based on the velocity of the 
impacting barge train, is displayed along the wall. The current location of 
the load is displayed as a red line along the time history. 

Because the impact can have a very low velocity, it can sometimes be hard 
to see the full impact in the wall view. For that reason, the time history 
display with a red line showing the current time is shown at the bottom left 
of the plot for the input impact-force time history. 

The rest of the area is the 2-D plot of the displacements of each node 
relative to the wall. The scaled displacements are shown in the scale along 
the X-dimension. The nodal displacements are connected with line 
segments, effectively linearly interpolating the displacements between 
nodes. 



ERDC/ITL TR-16-1 


132 


Figure 5.25 Animated graph of wall X-displacement. 


FEO Nodal Output 


File 

lmpact_Deck_T estProbleml .feo 


© Display extreme values and their times 

O Plot displacement vs. time Node Index; [to -r | Displacement/Rotation: [z~ 


® Plot animated MHliyifllHlNl 008400 g] Loop Displacement/Rotation: 


Time = 0.08400 X Displacement 




0.04 0.05 


0.06 0.0663 


Figure 5.26 Animated graph of wall Y-displacement. 


FEO Nodal Output 


lmpact_Deck_TestProbleml .feo 


© Display extreme values and their times 
O Plot displacement vs. time Node Index; [to 

® Plot animated IWIRTFimHIll • .08400 [gj| g] [joop Displacement/Rotation; 


Time = 0.08400 Y Displacement 


838.6667 

800.0 




0.0002 0.0005 0.0007 0.001 0.0012 0.0015 0.0017 

Oy (ft) 



















































ERDC/ITL TR-16-1 


133 



5.7.2 FEO Element Output 

Pressing the Show Element Information button from the Output tab 
brings up the FEO Element Output window. The window has a minimal 
menu, with only options to save the currently displayed information 
(either as text or as a .PNG bitmap graphic file for graphical data) and to 
close the window. The currently selected file name is displayed 
immediately below the menu. Beneath the menu are radio buttons that 
allow the user to present different information from the .FEO file. 

The FEO Nodal Output window presents element information in three 
ways 

1. The minimum and maximum values of all the elements are presented as 
text 

2. Individual elements are shown as a 2-D plot of force/moment versus time 

3. The entire structure is shown as a 2-D animated plot with the structure 
along the Y-axis, and the force/moment along the X-axis. 

When the “Display extreme values and their times” radio button is 
selected, a set of lists containing each element’s minimum and maximum 





























ERDC/ITL TR-16-1 


134 


force/moment (axial force, shear force, and moment) and the time that 
those extreme values occurred is displayed (Figure 5.28). 


When the “Plot force/moment vs. time” radio button is selected, the user 
can select an element to view using the “Element Index” combo box. When 
an element has been selected, the user can decide whether to plot the 
element’s Axial force, Shear force, or Moment versus Time in the plot 
below by selecting from the “Mode” combo box. 


Figure 5.28 FEO element output window showing maximum and minimum force and 
moments acting on all the elements. 


FEO Element Output 




File 





lmpact_Deck_TestProblem1.feo 




(«} Display extreme values and their times 



1 O Plot force/moment vs. time 

Element Index [ 

1 ▼ | Mode: 

| Axial ▼] 

O Plot animated [MINIMI! l[ Miff] 

Loop Mode: [Axial ▼] 


Axial 

Axial 

Axial 

Axial * 


Minimum 

Minimum 

Maximum 

Maximum 

Elem 

Force 

Time 

Force 

Time 

ID 

(kip) 

(sec) 

(kip) 

(sec) 

1 

-251.45 

0.21 

251.45 

0.21 

2 

-251.45 

0.21 

251.45 

0.21 

3 

-251.45 

0.21 

251.45 

0.21 

4 

-251.46 

0.21 

251.46 

0.21 

5 

-251.46 

0.21 

251.46 

0.21 

6 

-251.47 

0.21 

251.47 

0.21 

7 

-251.47 

0.21 

251.47 

0.21 

8 

-251.48 

0.21 

251.48 

0.21 

9 

-251.49 

0.21 

251.49 

0.21 

10 

-251.5 

0.21 

251.5 

0.21 

11 

-251.52 

0.21 

251.52 

0.21 

12 

-251.53 

0.21 

251.53 

0.21 

13 

-251.55 

0.21 

251.55 

0.21 

14 

-251.56 

0.21 

251.56 

0.21 

15 

-251.58 

0.21 

251.58 

0.21 

16 

-251.6 

0.21 

251.6 

0.21 

17 

-251.65 

0.21 

251.65 

0.21 

18 

-251.59 

0.21 

251.59 

0.21 

19 

-251.64 

0.21 

251.64 

0.21 

20 

-251.66 

0.21 

251.66 

0.21 

21 

-251.69 

0.21 

251.69 

0.21 

22 

-251.72 

0.21 

251.72 

0.21 

23 

-251.74 

0.21 

251.74 

0.21 

24 

-251.77 

0.21 

251.77 

0.21 

25 

-251.8 

0.21 

251.8 

0.21 

26 

-251.84 

0.21 

251.84 

0.21 

27 

-251.87 

0.21 

251.87 

0.21 




nr-, ft-. 

A A1 


When these options have been selected, the graphic window below shows 
the relevant plot (Figures 5.29 through 5.31). At the top of the display is 
the selected element number and selected force/moment. To the left of the 
plot is a display of the coordinate system and the wall with nodes and 
element connectivity. The selected element is displayed in red to highlight 
where the element is in relation to the wall. The rest of the area is the 2-D 
plot of the force/moment versus time for the selected element and 
force/moment. When the mouse is placed over this plot, a tooltip is shown 
revealing the force/moment for the time that the cursor is over. 























ERDC/ITL TR-16-1 


135 

















































ERDC/ITL TR-16-1 


136 


Figure 5.31 Graph showing element 60 shear force vs time. 



When the “Plot animated” radio button is selected, the user can select 
whether to plot the wall’s Axial force, Shear force, or Moment versus Time 
in the plot below by selecting from the “Mode” combo box. The 
force/moment at each element is linearly interpolated from element to 
element. The animation control behaves in the same manner as it did for 
the FEO Nodal Output window. 

When these options have been selected, the graphic window below shows 
the relevant plot (Figures 5.32 through 5.34). At the top of the display is 
the current time and selected force/moment mode. To the left of the plot is 
a display of the coordinate system and the wall with nodes and element 
connectivity. The plot of the impact-load time history from the starting 
position of the impact to the ending impact location, based on the velocity 
of the impacting barge train, is displayed along the wall. The current 
location of the load is displayed as a red line along the time history. 

Because the impact can have a very low velocity, it can sometimes be hard 
to see the full impact in the wall view. For that reason, the time history 
display with a red line showing the current time is shown at the bottom left 
of the plot for the input impact-force time history. 






















ERDC/ITL TR-16-1 


137 


Figure 5.32 Animated graph of axial forces acting on the wall. 


FEO Element Output 


lmpact_Deck_TestProblem1 feo 


© Display extreme values and their times 
© Plot force/moment vs. time Element Index; [70 
® P'otanimatad >■««» 


"BBiw 



0.0 

Axial Force (kips) 


Figure 5.33 Animated graph of moments acting on the wall. 


FEO Element Output 


| lmpact_Deck_Test Problem 1 .feo 


© Display extreme values and their times 

© Plot force/moment vs. time Element Index [l Mode: [Axial 

® Plotanimatad [MfflfMIIlHlll M ® “■» 


Mode: Moment ▼ 




0.0 

Mom ent (kip-ft) 


5000 £0662.18 







































ERDC/ITL TR-16-1 


138 


Figure 5.34 Animated graph of shears acting on the wall. 


FEO Element Output 


lmpact_Deck_TestProblem1 feo 


© Display extreme values and their times 

Plot force/moment vs. time Element Index: [ 70 ▼] Mode: [Moment 

® Plot animated WBSMMM [0.08400 [4-j| |7] Loop Mode: [shear 




-250.0 0.0 250.0 

Shear Force (kips) 


The rest of the area is the 2-D plot of the force/moment of each element 
relative to the wall. The scaled force/moment is shown in the scale along 
the X-dimension. The element force/moment is projected to the nodes and 
the values connected with line segments, effectively linearly interpolating 
the force/moments between elements. 

5.7.3 FEO Pile Group Response 

Pressing the Show Pile Group Response button from the Output tab brings 
up the Pile Group Response window. The window has a minimal menu, 
with only options to save the currently displayed information (either as 
text or as a .PNG bitmap graphic file for graphical data) and to close the 
window. The currently selected file name is displayed immediately below 
the menu. Beneath the menu are radio buttons that allow the user to 
present different information from the .FEO file. 

The Pile Group Response window presents element information in three 
ways: 

l. The minimum and maximum values of all the pile group nodes are 
presented as text 



















ERDC/ITL TR-16-1 


139 


2. The force and moment values at any selected time during the analysis are 
presented as text 

3. The entire structure is shown as a 2-D animated plot with the structure 
along the Y-axis, and the pile group force/moment along the X-axis. 

When the “Display extreme values and their times” radio button is selected, 
a list containing each pile group nodes’s minimum and maximum group 
response force/moment (longitudinal force, transverse force, and moment) 
and the time that those extreme values occurred (Figure 5.35) and peak 
displacements and when they occurred (Figure 5.36) are displayed. 


When the “Display Spring Forces at Time” radio button is selected, the 
user can enter a time in the simulation and get a snapshot of the forces 
acting at each pile group node at that time (Figure 5.37). At the base of the 
list, the forces are totaled to give an overall force acting at all of the pile 
group nodes. Additionally, the total impulse calculations for longitudinal 
and transverse forces and the moments are presented at the end of the list. 


Figure 5.35 Pile Group Response Maximum and Minimum Forces and Moments. 


Pile Group Response 


lmpact_Deck_TestProbleml feo 


1 <§> Displa 
I O Displa 
1 O PlotAi 

ly extreme values and their times 

iy Spring Forces at Time: 0.0000 


Loop Pile Grou| 




limated M Ml M W POOOO 

p Node Index [l 

▼ | Degree of Freedom: (x ▼] 







Long. 

Long. 

Trans. 

Trans. 



> 

Node 

Force 

Time 

Force 

Time 

Moment 

Time 


ID 

(kips) 

(sec) 

(kips) 

(sec) 

(kip-ft) 

(sec) 

E 

1 

251.4489 

0.21 

0.9333 

0.142 

0.0012 

0.0012 

J 

2 

0.0047 

0.21 

0.001 

0.18 

0 

0 

3 

0.0138 

0.21 

0.0028 

0.18 

0 

0 

4 

0.0229 

0.21 

0.0045 

0.18 

0 

0 

5 

0.032 

0.21 

0.006 

0.18 

0 

0 

6 

0.0411 

0.21 

0.0073 

0.178 

0 

0 

7 

0.0502 

0.21 

0.0083 

0.178 

0 

0 

8 

0.0593 

0.21 

0.009 

0.178 

0 

0 

9 

0.0684 

0.21 

0.0093 

0.178 

0 

0 

10 

0.0775 

0.21 

0.0093 

0.178 

0 

0 

11 

0.0866 

0.21 

0.0089 

0.176 

0 

0 

12 

0.0957 

0.21 

0.0082 

0.176 

0 

0 

13 

0.1048 

0.21 

0.0078 

0.086 

0 

0 

14 

0.1139 

0.21 

0.0082 

0.082 

0 

0 

15 

0.123 

0.21 

0.0088 

0.08 

0 

0 

16 

0.1321 

0.21 

0.0096 

0.078 

0 

0 

17 

0.1412 

0.21 

0.0105 

0.076 

0 

0 

19 

0.1505 

0.21 

0.0092 

0.076 

0 

0 

20 

0.1596 

0.21 

0.0059 

0.072 

0 

0 

21 

0.1687 

0.21 

0.0059 

0.14 

0 

0 

22 

0.1778 

0.21 

0.007 

0.134 

0 

0 

23 

0.1869 

0.21 

0.0087 

0.13 

0 

0 

24 

0.196 

0.21 

0.0105 

0.126 

0 

0 

25 

0.2051 

0.21 

0.0121 

0.124 

0 

0 

26 

0.2143 

0.21 

0.0134 

0.124 

0 

0 

27 

0.2234 

0.21 

0.0144 

0.122 

0 

0 

28 

0.2325 

0.21 

0.0148 

0.12 

0 

0 

29 

0.2416 

0.21 

0.0149 

0.118 

0 

0 

- 


































ERDC/ITL TR-16-1 


140 


Figure 5.36 Pile Group Response Peak Deflections. 



Figure 5.37 Pile Group Response Maximum and Minimum Forces and Moments. 














































ERDC/ITL TR-16-1 


141 


When the “Plot Animated” radio button is selected, the user can plot an 
individual pile group node’s force-versus-displacement over time. Using 
the “Pile Group Node Index” combo box, the user can select a pile group 
node for display. The “degree of freedom” combo box provides the axis 
that the resulting force is applied. The animation control behaves in the 
same manner as it did for the FEO Nodal Output window. 

When these options have been selected, the graphic window below shows 
the relevant plot (Figures 5.38 through 5.40). At the top of the display is 
the current time, the selected pile group node, and selected force axis. To 
the left of the plot is a display of the coordinate system and the wall with 
nodes and element connectivity. The selected pile group node is 
highlighted in red to show the pile group node position relative to the 
structure. The plot of the impact-load time history from the starting 
position of the impact to the ending impact location, based on the velocity 
of the impacting barge train, is displayed along the wall. The current 
location of the load is displayed as a red line along the time history. 


Figure 5.38 Animated plot of node 85 X-force and displacement vs time. 
























ERDC/ITL TR-16-1 


142 


Figure 5.39 Animated plot of node 85 Y-force and displacement vs time. 


Pile Group Response 


lmpact_Deck_TestProbleml .feo 


O Display extreme values and their times 


© Display Spring Forces at Time: 3 2600 


Plot Animated MHTHIUMliHl 0.26000 _ 


|4J| g] Loop Pile Group Node Index: 185 ▼] Degree of Freedom: hr 


Time = 0.26000 Node: 85 Y-Axis 



Figure 5.40 Animated plot of node 85 Z-force and displacement vs time. 

Pile Group Response 


File 

lmpact_Deck_TestProbleml feo 


1 O Display extreme values and their times 

1 O Display Spring Forces at Time: 0.2600 


1 ® Plo, Animated | IjlNll^l Mil MlW] 0.26000 

g] Loop Pile Group Node Index: [85 ▼ ] Degree of Freedom: [z ▼] 


Time = 0.26000 Node: 85 Z-Rotation 


0.25 

0.2 


Force (kips) 


0.05 

0.04 

0.03 

0.02 

0.01 

0 . 0 - 

- 0.01 

- 0.02 

-0.03 

-0.04 

-o.g 


Force (kips) 


- 0.1 


0.005 -0.0025 0.0 0.0025 0.005 

Displ. (in) 


Displ. (in) 


- 0.2 

- °- 2 5.0 1.0 2.0 3.0 4.0 5.0 

Time (sec) 


0.25 

0.2 1 

0.1 

0 . 0 - 

- 0.1 


- 0.2 

~°- 2 §.0 1.0 2.0 3.0 4.0 5.0 

Time (sec) 



































ERDC/ITL TR-16-1 


143 


Besides the wall plot, there are three additional plots. The plot in the 
upper left is the resultant force-versus-time plot. This force results from 
spring deformation along the selected axis. A red point shows the current 
time and its force for the selected pile group node. 

The plot in the lower left is the displacement versus time plot. A red point 
shows the current time and its displacement for the selected pile group 
node. 

The center plot shows the force-versus-displacement plot. This plot shows 
how force varies with displacement at the pile group node. Because this 
graph has axis with values that may be repeated (and therefore is not a 
mathematical function), the current force and time are located with cross¬ 
hairs for the current time. 

Note that the Z-axis DOF is not a 3-D axis but a rotational axis at the pile. 
For pile group nodes that are not connected due to a moment release 
between deck sections, the force versus the rotational displacement will 
give a zero value. 

5.7.4 Run Information 

Pressing the Show Run Output button from the Output Tab brings up a 
text window with the results of the program run (Figure 5.41). The menu 
allows the user to save the file to a new location or exit the Run Output 
window. How the program worked is displayed in this window, as well as 
any specific finite element data requested in the Analyze Tab. 



ERDC/ITL TR-16-1 


144 


Figure 5.41 Output run information with selected FEO node output. 


[mpact_Deck_TestProbleml.out - — 4 



File 

Maximum applied 

normal force 

_ 

516.442073 


Maximum applied 

tangent force = 

258.221037 


Time for maximum applied forces = 

0.200000 



Located at x = 

501.389500 


Spring "Y" 

Maximum 

Time max. 

Maximum 

Time max. 

at Node Displacement 

Displ. 

Force 

Force 

2 

0.0000 

0.2100 

0.0047 

0.2100 

3 

0.0000 

0.2100 

0.0138 

0.2100 

4 

0.0001 

0.2100 

0.0229 

0.2100 

5 

0.0001 

0.2100 

0.0320 

0.2100 

6 

0.0001 

0.2100 

0.0411 

0.2100 

7 

0.0001 

0.2100 

0.0502 

0.2100 

8 

0.0002 

0.2100 

0.0593 

0.2100 

9 

0.0002 

0.2100 

0.0684 

0.2100 

10 

0.0002 

0.2100 

0.0775 

0.2100 

11 

0.0002 

0.2100 

0.0866 

0.2100 

12 

0.0003 

0.2100 

0.0957 

0.2100 

13 

0.0003 

0.2100 

0.1048 

0.2100 

14 

0.0003 

0.2100 

0.1139 

0.2100 

15 

0.0003 

0.2100 

0.1230 

0.2100 

16 

0.0004 

0.2100 

0.1321 

0.2100 

17 

0.0004 

0.2100 

0.1412 

0.2100 

19 

0.0004 

0.2100 

0.1505 

0.2100 

20 

0.0005 

0.2100 

0.1596 

0.2100 

21 

0.0005 

0.2100 

0.1687 

0.2100 

22 

0.0005 

0.2100 

0.1778 

0.2100 

23 

0.0005 

0.2100 

0.1869 

0.2100 

24 

0.0006 

0.2100 

0.1960 

0.2100 

25 

0.0006 

0.2100 

0.2051 

0.2100 

26 

0.0006 

0.2100 

0.2143 

0.2100 

27 

0.0006 

0.2100 

0.2234 

0.2100 

28 

0.0007 

0.2100 

0.2325 

0.2100 

29 

0.0007 

0.2100 

0.2416 

0.2100 

30 

0.0007 

0.2100 

0.2507 

0.2100 

31 

0.0007 

0.2100 

0.2598 

0.2100 

32 

0.0008 

0.2100 

0.2689 

0.2100 


□ 


5.8 Example: Geometry Input for the Impact Deck at Lock and 
Dam 3 

The example problem for the Lock and Dam 3 impact deck that was input 
in section 2 present the most complex geometry for GUI input. In this 
case, the approach wall structure had 8 reinforced concrete monoliths. 
These monoliths were connected to each other and with the single end cell 
at the start of the approach wall with a pinned connection, and to the 
neighboring monoliths through a connection that transfer shear and axial 
forces but not moments. 


The monoliths themselves were to be constructed by connecting together 
8 concrete segments that were 12 ft 6 in. in length. Each segment was 
supported by two pile groups. Each pile group cluster consisted of 3 piles - 
a vertical pile at the front of the group, closest to where an impact would 
occur, and two batter piles with a batter of 1:4. The spacing between the pile 
groups was to be 6 ft 3 in., which resulted in monoliths with a length of 
100 ft. These plans changed, as engineering plans often do, due to wall 
length requirements. The new monolith length ended up becoming 104 ft 
10 in. The distance between pile groups grew to 6 ft 3.5 in. The distance 
from the ends of the monolith to the first and last pile group was 3 ft 4.25 in. 







ERDC/ITL TR-16-1 


145 


In this case, it was natural to model each monolith as a set of nodes in the 
Impact_Deck GUI, and then copy the monolith nodes 7 times (for 8 
monoliths). The end nodes of the first monolith were the connecting nodes. 
These nodes were at o ft o in. and 104 ft 10 in., which translated to decimal 
feet of 0.0 ft and 104.833 ft. Because inches did not convert to nice decimal 
feet, the user must determine his or her personal level of precision. Because 
that last node was coincident with the first node of the next monolith, the 
last node does not needed to be entered as it will be created in the copy and 
offset command. Therefore, the user only needs to place a node at position 
0.0 ft. Because it was an inter-monolith node, the check-box was checked to 
flag the node (Figure 5.42). 

The rest of the nodes in the monolith represented each pile group location 
in the monolith. These monolith groups started 3 ft 4.25 in. from the start 
of the monolith (at o ft o in.) and ended at 3 ft 4.25 in. from the end of the 
monolith (at 104 ft 10 in.). These coordinates in decimal feet were 3.354 ft 
and 101.479 ft, respectively. There were 16 pile groups in the monolith 
with 15 divisions between them. Entering these values into the 
“Interpolated Node Input” box (with the inter-monolith nodes check-box 
unchecked) created the nodes for the pile groups (Figure 5.43). 

Figure 5.42 Add node at position 0.0 ft as an inter-monolith node. 







































ERDC/ITL TR-16-1 


146 


Figure 5.43 Add interpolated nodes from 3.3541666 ft to 101.4791666 ft. 



When the set of nodes for the monolith were created (without the ending 
inter-monolith nodes), all of the nodes for the monolith were selected using 
a left-mouse, click-drag operation (Figures 5.44 and 5.45). Clicking the 
“Copy Selected Nodes...” button allowed the user to create copies of this 
monolith node set multiple times (Figure 5.46). Enter in the offset of the 
monolith as 104 ft 10 in. (104.833 decimal feet- Figure 5.47), then click the 
OK button in the “Copy Selected Nodes with Offset” window seven times 
(Figure 5.48). 


Figure 5.44 Selecting nodes with a left-mouse, click-drag. 



Figure 5.45 Selected nodes are shown with vertical lines. 


Input Plot 


r Y 

x 


1 


o Pile Group Node 


Node 16 
YPos: 

94.9375000003 

333 























































































ERDC/ITL TR-16-1 


147 


Figure 5.46 Selected nodes can be copied multiple times with the copy selected 

nodes button. 



Figure 5.47 The copy selected nodes dialog lets the user specify 

an offset. 



Figure 5.48 Select OK the number of times that the 
selected nodes need to be copied. 









































































































ERDC/ITL TR-16-1 


148 


Because the monolith was modeled without the ending inter-monolith 
connection, the last connection was added using the “Single Node Input” 
box. Its position was at 838’-8” (838.666 decimal feet). This node was 
created as an inter-monolith node (Figure 5.49). 

Figure 5.49 Finally, Add the Final Node. 



Hopefully, this subsection has revealed the usefulness of the Impact_Deck 
GUI methods for modeling geometry using interpolation and copying of 
repetitive structures that might otherwise require a good amount of 
calculation for the user. 

5.9 Final Remarks 

In this section, the user was presented with the GUI specifications. From 
this information, the user should be able to define the geometry, select a 
force time history for an impact, give beam and pile group properties, and 
perform an analysis from this input model. After the analysis was 
performed, methods for varying visualization of results were presented, 
either as static or dynamic plots with information for nodes, elements, and 
resultant forces at pile groups. The user was also presented with examples 
that show input for the various structures, and the GUI output (impact 
decks, flexible approach walls, and guard walls). 
























































ERDC/ITL TR-16-1 


149 


6 Conclusions and Recommendations 

In this research project, an involved dynamic time-domain analysis was 
performed to determine the displacement and response forces of flexible 
pile groups supporting a concrete beam or deck used for the absorption of 
energy from a barge train impact. An impact force time history was used to 
represent an impact event (Ebeling et al. 2010). Three different types of 
pile-founded, flexible approach walls were studied: an impact deck, an 
alternative flexible approach wall, and a guard wall. 

The first case was the Lock and Dam 3 impact deck structure consisting of 
eight concrete monoliths. These monoliths were supported over a series of 
equally spaced rows of three cluster pile groups. An internal pin (i.e., with 
no bending moment transfer) formed the connection between adjacent 
monoliths. At one end of the structural system, the impact deck monolith 
was pin-connected to a massive concrete circular cell and at the other end 
of the structural system the impact deck monolith was free. This structural 
system was modeled by means of typical beam elements between each 
pile-group row or between the last pile group row of the monolith to the 
inter-monolith pin connection. Each pile-group was modeled by using a 
pair of elastic-plastic translational springs. The definitions of the spring 
stiffness was determined by doing a push-over analysis of a single pile- 
group cluster. The dynamic-impact load had a specified starting point and 
was stationary or moving along the wall. A damping effect was included 
using the Rayleigh damping theory, which depends on the natural 
frequency of the system. The two natural frequencies of the structural 
system (to be used for Rayleigh damping) were calculated in an 
approximate form by using the spring stiffness and the mass of the impact 
beam or deck. After the global mass, damping, and stiffness matrices and 
the load vector were assembled, the dynamic time-history response of the 
system was calculated. 

The second case was the McAlpine alternative flexible approach wall. A 
section of the approach wall consisted of two consecutive concrete beams 
supported over three pile groups was modeled with the software. The 
effect of non-impacted beams was modeled simplistically and the user 
specifies how many beams before and after the area of interest exist (i.e., 
where the impact event occurs). The connection of the first beam to the 
second beam at the center pile cap was accomplished by using a shear key 
(i.e., no bending moment transfer). In fact, due to the distance between 



ERDC/ITL TR-16-1 


150 


the end of one beam and the start of the second beam, the shear key had to 
be modeled also as two rigid beam elements, which connected at a node 
centered between the two beam end nodes. From there, a rigid beam 
element was perpendicularly connected back to a node at the center of 
rigidity of the pile group and its cap. This established the position of the 
two translational springs and one rotational spring support. In this way, 
the force-translational and moment-rotational effects of the central pile 
group were modeled. This structural system was modeled by means of 
typical beam elements between each pile group and the three rigid beam 
elements of the central pile cap. The pile groups at the start and end of the 
structural system were modeled by using elastic-plastic translational 
springs. The central pile group was modeled by using two elastic-plastic 
translational springs and one elastic-plastic rotational spring. The 
definitions of the translational spring stiffness was determined by doing a 
push-over analysis of a typical pile group. The rotational spring properties 
were defined by using the translational spring properties of the pile group 
as shown in Appendix E. The dynamic impact-force time history had a 
specified starting point and was stationary or moving along the beam. A 
damping effect was included using the Rayleigh damping theory, which 
depended on the natural frequency of the system. The two natural 
frequencies of the structural system (to be used for Rayleigh damping) 
were calculated in an approximate form by using the spring stiffness and 
the mass of the beam. After the global mass, damping, and stiffness 
matrices and the load vector were assembled, the dynamic time-history 
response of the system was calculated. 

The third case was a typical guard wall. A section of the approach wall 
consisted of two consecutive concrete beams supported over three pile 
groups was modeled with the software. The connection of the first beam to 
the second beam in the center pile cap was achieved by using a shear key 
(i.e., no bending moment transfer). This structural system was modeled by 
means of typical beam elements between each pile row. The beam 
elements that connect at the central pile row were modeled with end 
releases (i.e., no moment transfer). The start, central, and end pile rows of 
the structural system were modeled by using a pair of elastic-plastic 
springs. The definitions of the stiffness for the translational and 
longitudinal springs were determined by doing a push-over analysis of a 
typical pile group. The dynamic impact load had a specified starting point 
and was stationary or moving along the beam. A damping effect was 
included using the Rayleigh damping theory, which depended on the 



ERDC/ITL TR-16-1 


151 


natural frequency of the system. The two natural frequencies of the 
structural system (to be used for Rayleigh damping) were calculated in an 
approximate form by using the spring stiffness and the mass of the impact 
beam or deck. After the global mass, damping, and stiffness matrices and 
the load vector were assembled, the dynamic response of the system was 
calculated. 

The general input data for these three models were the nodes and their 
positions along the wall, modulus of elasticity of the beam, cross-sectional 
area of the beam, moment of inertia of the beam, mass per unit length of 
the beam, damping ratio, dynamic coefficient of friction between the lead 
impact barge and the wall, the initial point of contact, velocity of the 
moving load, springs properties, and force time history description. The 
accuracy of the solution depended on the appropriateness of these 
variables. 

To demonstrate the effectiveness of the methodology developed in the 
Impact_Deck computer program, a validation against SAP2000 and 
several examples were presented. The validation produced outstanding 
results. In the examples when the plastic behavior was reached, the time 
history results were in agreement with the developed elastic-plastic, force- 
displacement relationship of the springs. It was recognized that SAP2000 
will not handle a moving dynamic load nor the nonlinear springs used to 
model the unique clustered pile groups responses. These are two unique 
capabilities of the Impact_Deck software. 

An important conclusion from the Impact_Deck analyses was that inertial 
effects of the wall superstructure and substructure during dynamic loading 
were important to the computed results and should not be ignored. A 
dynamic analysis must be performed because the resulting overall peak 
response force was much greater than the peak input impact force and the 
overall peak response force happened much later in time than when the 
peak input impact force occured. For every structure analyzed, the overall 
peak response force had a greater value than the peak input impact force 
value and the time to the overall peak response force was greater than the 
time to the peak input impact force. For the simply supported beam guard 
wall model (in section 4) the overall response force was 181% of the peak 
input impact force. The overall peak response time was 0.392 sec, or 
0.192 sec after the peak input impact force time of 0.2 sec. 



ERDC/ITL TR-16-1 


152 


The peak response force for any single pile-group node as a substructure 
for a simply supported beam, in situations where load sharing was 
minimal (in the McAlpine alternative flexible approach wall and guard 
wall examples), also exceeded the peak input load force and, due to inertial 
effects, happened at a later time than the peak input impact force. For the 
McAlpine alternative flexible approach wall and the guard wall examples, 
the peak response force for the single pile group node with the greatest 
force was and 123% and 128% of the peak input load force, respectively. 
These peaks occurred 0.08 and 0.132 sec after the time to peak input 
impact force. These were important considerations for the design of pile 
supported approach wall structures subjected to barge train impact 
loading. 

The peak response force for any single pile-group node with a fixed 
connection to a deck, and where multiple pile groups support the monolith 
deck, benefited from the effect of load sharing. In this case, the motion of 
the impact deck generated a response from all of the pile groups (16, in our 
example). Because of the shear key connection between monoliths, the pile 
groups of all 8 monoliths responded similarly to the impact deck motion. 
However, despite load sharing at 0.04 sec after the peak input load, the 
peak response force for any pile group reached its peak value of 56.456 
kips. This was 11% of the peak input load force, or 7.8% of the peak overall 
pile-group response force. While it was reasonable to assume a force 
reduction of greater than 16 times could occur, the real reduction was only 
slightly more than 9 times less for the peak input load and nearly 13 times 
less for the peak overall pile group response. Load sharing occured, but its 
effects were not as pronounced, due to the inertia of the impact deck under 
a dynamic barge train time history load. These results demonstrated the 
advantage of using this moment resistant impact deck monolith supported 
on a large number of smaller piles versus using a simply supported impact 
beam with larger piles and longer spans: the design forces for each 
individual pile group was much less. Site conditions and the use of clever, 
cost-effective, in-the-wet and above-the-wet construction practices 
(including the use of precast structural features) will dictate which design 
type will ultimately possess the greater advantage (in cost and effort) for 
an approach wall project. 



ERDC/ITL TR-16-1 


153 


References 


Bathe, K. 1996. Finite element procedures, Englewood Cliffs, New Jersey: Prentice Hall. 

Chopra, Anil K. 1995. Dynamics of structures; theory and applications to earthquake 
engineering, Englewood Cliffs, New Jersey: Prentice Hall. 

Chopra, Anil K. 2001. Dynamics of structures; theory and applications to earthquake 
engineering, second edition, Englewood Cliffs, New Jersey: Prentice Hall. 

Clough, Ray W., and Joseph Penzien. 1993. Dynamics of structures, Second edition, New 
York, NY: McGraw-Hill, Inc. 

Computers and Structures, Inc. 2003. SAP2000 static and dynamic finite element 
analysis of structures nonlinear 8.2.3, Berkeley, CA, 2003. 

Craig, R. R. 1981. Structural dynamics, an introduction to computer methods, New York, 
NY: John Wiley & Sons, Inc. 

Davisson, M. T. 1970. Lateral load capacity of piles, Highway Research Record, Number 
133, Pile4 Foundations, Washington, D.C.: Highway Research Board. 

Ebeling, Robert M., R. A. Green, and S. E. French. 199 y. Accuracy of response of single 

degree-of-freedom systems to ground motion, Earthquake Engineering Research 
Program, TRITL-97-7, Vicksburg, MS: U.S. Army Waterways Experiment 
Station. 

Ebeling, Robert M., Barry C. White, Abdul N. Mohamed, and Bruce C. Barker. 2010. 

Force time-history during the impact of a barge train impact with a approach 
lock wall using impact^force, ERDC/ITL TR-10-1, Vicksburg, MS: U.S. Army 
Engineer Research and Development Center. 

Ebeling, Robert M., Abdul N. Mohamed, Jose R. Arroyo, Barry C. White, Ralph W. Strom, 
and Bruce C. Barker. 2011. Dynamic structural flexible-beam response to a 
moving barge train impact force time-history using impact_beam, ERDC/ITL 
TR-11-1, Vicksburg, MS: U.S. Army Engineer Research and Development Center. 

Ebeling, R. M., R. W. Strom, B. C. White, and K. Abraham. 2012. Simplified Analysis 

Procedures for Flexible Approach Wall Systems Founded on Groups of Piles and 
Subjected to Barge Train Impact, ERDC/ITL TR-12-3, Vicksburg, MS: U.S. 
Department of the Army, Army Corps of Engineers, Engineer Research and 
Development Center. 

Hartman, J. P., Jaeger, J. J., Jobst, J. J., and Martin, D. K. 1989. User's Guide: Pile 

Group Analysis (CPGA). Technical Report ITL-89-3. Vicksburg, MS: U.S. Army 
Engineer Waterways Experiment Station. 

Hilber, H. M., T. J. R. Hughes, and R. L. Taylor. 1977. Improved Dissipation for Time 
Integration Algorithms in Structural Dynamics. Earthquake Engineering and 
Structural Design. 5:283-292. 

MathCAD 8.1998. Cambridge, MA: Mathsoft, Inc. 



ERDC/ITL TR-16-1 


154 


McGuire, W., and R. H. Gallagher. 1979. Matrix structural analysis, New York, NY: John 
Wiley & Sons, Inc. 

Patev, Robert C., Bruce C. Barker, and Leo V. Koestler. 2003. Prototype barge impact 
experiments, Allegheny lock and dam 2, Pittsburgh, Pennsylvania, ERDC/ITL 
TR-03-2, Vicksburg, MS: U.S. Army Engineer Research and Development Center. 

Paz, Mario. 1985. Structural dynamics; theory and computation, Second edition, New 
York, NY: Van Nostrand Reinhold Company. 

Paz, Mario. 1991. Structural dynamics; theory and computation, Third edition, New 
York, NY: Van Nostrand Reinhold Company. 

Press, W. H., S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery. 1996. Numerical 

recipes in Fortran 77, the art of scientific computing, Second edition, New York, 
NY. 

Reddy, J. N. 1993. An introduction to the finite element method, Second edition, 
McGraw-Hill, Inc. 

Rangan, B. V., and M. Joyce. 1992. Strength of eccentrically loaded slender steel tubular 
columns filled with high-strength concrete, ACI Structural Journal, 89(6): 676- 
681. 

Ross, C. T. F. 1991. Finite element programs for structural vibrations, New York, NY: 
Springer-Verlag Berlin Heidelberg. 

Saul, W. E. 1968. Static and dynamic analysis of pile foundations, Journal of the 

Structural Division, ASCE, Volume 94, Number St5, Proceeding Paper 5936. 

Stadler, W., and R. W. Shreeves. 1970. The transient and steady-state response of the 

infinite Bernoulli-Euler beam with damping and an elastic foundation. Quarterly 
Journal of Mechanics and Applied Mathematics, 23(2): 197-208. 

Tedesco, J., W. G. McDougal, and C. A. Ross. 1999. Structural dynamics-theory and 
applications, Menlo Park, California: Addison Wesley Longman, Inc. 

Terzaghi, K. 1955. Evaluation of coefficient of subgrade reaction, Geotechnique, Vol. 5, 
pp. 297-326. 

Yang, N. C. 1966. Buckling strength of pile, Highway Research Record, Number 147, 
Bridges and Structures, Washington, D.C.: Highway Research Board. 

Wilson, E. L. 2002. Three-dimensional static and dynamic analysis of structures-A 
physical approach with emphasis on earthquake engineering, Berkeley, 

California: Computer and Structures, Inc. 

Wilson, E. L. 2010. Static & Dynamic Analysis of Structures: A Physical Approach with 
Emphasis on Earthquake Engineering, 4 th edition, p. 394, Berkeley, CA: 
Computers and Structures, Inc. 

Weaver, W., and P. Johnston. 1984. Finite elements for structural analysis, Englewood 
Cliffs, New Jersey: Prentice Hall. 



ERDC/ITL TR-16-1 


155 


Appendix A: Lock and Dam 3 - Equations of 
Motion for the Mathematical Model 

In structural dynamics, the mathematical model of bodies of finite 
dimensions undergoing translational motion are governed by Newton’s 
Second Law of Motion, expressed as 

^F = m«a (1.1 bis) 

at each time step t during motion. In the mathematical model of 
transverse vibration, the forces acting on the flexible impact beam mass at 
each time step t include (l) the impact force at time step t, (2) the elastic 
restoring forces (of the beam), and (3) the damping forces (of the beam). 
The mathematical model of the beam in the engineering formulation 
described in this appendix has a finite number of degrees of freedom 
(DOF) because it is discretized using the finite element formulation. The 
engineering formulations of equations of motion are solved using a 
numerical solution method to determine the displacement and response 
forces at each pile bent support feature. Each group of clustered piles is 
modeled as transverse and longitudinal elastic/plastic springs. This is a 
dynamic process, with the response forces responding to an impact pulse 
force time history and each calculation of the equations of motion 
occurring at each time step. The applied impact force can be given a 
constant velocity parallel to the approach wall so that it is changing 
position with time. The responses of the elastic beam over elastic or plastic 
springs can be obtained by the use of the Multi-Degrees-of-Freedom 
model (MDOF) and the finite element formulation of the beam element. 
The Impact_Deck software includes damping forces by using the Rayleigh 
damping model. The response is obtained using either the HHT-a method 
(Appendix F) or the well-known Wilson -0 method (Appendix G). By using 
the Newton’s Second law and applying it to a MDOF system, the resulting 
equations of motion can be expressed as, 

[Af]{ii(t)} + [C]{li (t)} + [*]{u(t)} = (F(t)} (2.1 bis) 


where, 

[M] = global mass matrix 
[C] = global damping matrix 



ERDC/ITL TR-16-1 


156 


[K] = global stiffness matrix 
(F(t)} = global vector of external forces and moments 
(u(t)}, (u(t)}, {u(t)} = relative acceleration, relative velocity, and relative 
displacement for each DOF. 

This appendix discusses the various relationships used in this engineering 
methodology that are implemented in Impact_Deck. 

A.l Element degrees of freedom (DOF) and interpolation functions 

Consider a straight-beam element of length L, mass per unit length m(x), 
and flexural rigidity El (x). The two nodes by which the 2-D finite element 
can be assembled into a structure are located at its ends. If only planar 
displacements are considered, each node has three DOF: the longitudinal 
displacement, the transverse displacement, and rotation. 

The longitudinal displacement (i.e., axial direction) of the beam element is 
related to its two DOF: 


U ( X > t ) = Yj U i(. t W X ) (A- 1 ) 

1=1 

where the function <fii(x) defines the displacement of the element due to 
unit displacement u it while constraining other DOF to zero. Thus (pi(x) 
satisfies the following boundary conditions and are shown in Figure A.i, 


/ = 1:%(0) = 1, tp ± (L) = 0 

(A. 2 ) 

/ = 2 :\p 2 ( 0 ) = 0 , W 2 (L)= 1 

(A. 3 ) 


Note that these DOF subscript values i = l and 2 correspond to Figure A.i 
DOF subscripts i = 1 and 4. 

The transverse displacement and rotation of the beam element is related 
to its four DOF, 


u(x,t) = Yjti(t)Wi(x) i = 2,3,5,6 


(A.4) 



ERDC/ITL TR-16-1 


157 


where the function ify(x) defines the displacement of the element due to 
unit displacement u i: while constraining other DOF to zero. Thus, ify(x) 
satisfies the following boundary conditions 


Figure A.1 Shape function for axial displacement 
effect. 



i = 1: iM0)=l, (0) =% (L) =Wi (L) = 0 (A.5) 

i = 2 : ip'fOj = 1, ip 2 (0) =ip 2 (L) = \p 2 (L) = 0 (A.6) 

/ = 3 : ip 3 (. L ) = l,ip 3 (0) = ip 3 (0) = ip 3 (L) = 0 (A.7) 

i = 4:ip\(L) = l,ip 4 (0)=ip' 4 (0)=ip 4 (L) = 0 (A.8) 


Note that these DOF subscript values i = l, 2, 3 and 4 correspond to 
Figure A.2 DOF subscripts i = 2, 3, 5, and 6. 

These interpolation functions could be any arbitrary shapes satisfying the 
boundary conditions. One possibility is the exact deflected shapes of the 
beam element due to the imposed boundary conditions, but these are 
difficult to determine if the flexural rigidity varies over the length of the 
element. However, they can conveniently be obtained for a uniform beam as 
illustrated next for the transverse displacement and rotation. Neglecting 
shear deformations, the equilibrium equation for a beam loaded only at its 
ends is 











ERDC/ITL TR-16-1 


158 


„ T d A u . 

£/ d7 = ° (A ‘ 9) 

The general solution of Equation (A.9) for a uniform beam is a cubic 
polynomial 


u 


M 


V 

+ flo 

/ \ 

X 

2 

+ a a 

'x' 


0 

L, 


X, 


(A.10) 


The constants a t can be determined for each of the four sets of boundary 
conditions of Equations (A.5) to (A.8), to obtain 


Ti(*) = l - 3 


\y 2 (x) = L 


( ^ 

1 X 


+ 2 


( \3 

1 X ' 



( \ 

V 


( \ 

X 

2 

' x' 

L 

A 

-2 L 

+ L 


~L, 


L, 

L 


/ \ 

X 

2 

9 

/ \ 

X 


— z 

L 


T 3 W = 3 

W 4 (x) = -L(x/Lf+L(x/Lf 


(A.ll) 

(A.12) 

(A.13) 

(A.14) 


These interpolation functions, illustrated in Figure A.2, can be used in 
formulating the element matrices. The same process can be done with the 
axial effect to obtain the basic interpolation functions, 



(A.15) 

M> 2 ( x ) = § 

(A.16) 


The finite element method is based on assumed relationships between the 
displacements at interior points of the element and the displacements at 
the nodes. Proceeding in this manner makes the problem tractable but 
introduces approximations in the solution. 



ERDC/ITL TR-16-1 


159 


Figure A.2 Shape function for transverse 
displacement and rotation effect. 




X 


L 




O 




A.2 Element stiffness matrix 

Consider a beam element of length L with flexural rigidity EI(x). By 
definition, the stiffness influence coefficient /q ; - of the beam element is the 

force in the DOF i due to unit displacement in DOFj. Using the principle 
of virtual displacement, the general equation for k t j, which is the stiffness 

term for the transverse displacement and rotation, in the element stiffness 
matrix is 


L 



(A. 17 ) 


o 


The symmetric form of this equation shows that the element stiffness 
matrix is symmetric, k t j = kj t . Equation (A.17) is a general result in the 

sense that it is applicable to elements with arbitrary variation of flexural 
rigidity EI(x), although the interpolation functions of Equations (A.11) to 
(A.14) are exact only for uniform elements. The associated errors can be 
reduced to any desired degree by reducing the element size and increasing 
the number of finite elements in the structural idealization. For a uniform 
finite element with EI(x) = El, the integral of Equation (A.17) can be 
evaluated analytically for i,j = 2,3,5, and 6, resulting in the 
corresponding terms in the element stiffness matrix. 


















ERDC/ITL TR-16-1 


160 


In the same way, for the axial contribution, the general equation for /qy, 
which is the stiffness term for the axial displacement in the element 
stiffness matrix is 


L, 

ky = fEA(x)\p](x)\p'j(x)dx 


(A.18) 


For a uniform finite element with EA(x) = EA, the integral of Equation 
(A.18) can be evaluated analytically for i,j = 1,4 resulting in the 
corresponding terms in the element stiffness matrix. Finally, when all 
terms are calculated, the element stiffness matrix can be obtained as, 


K\ = 


AE 

L 

0 

0 

AE 

~L 

0 


0 


0 

0 

AE 

0 

0 

L 

Y 1 EI 

6EI 

0 

12 EI 

6EI 

L 3 

L 2 

L 3 

L 2 

6EI 

L 2 

4 El 

L 

0 

6EI 

L 2 

2 EI 

L 

0 

0 

AE 

L 

0 

0 

12 EI 

6EI 

0 

12 EI 

6 El 

L 3 

L 2 

L 3 

L 2 

6EI 

2 EI 

0 

6EI 

4 EI 

L 2 

L 

L 2 

L 


(A.19) 


These stiffness coefficients are the exact values for a uniform beam, 
neglecting shear deformation, because the interpolation functions are the 
true deflection shapes for this case. Observe that the stiffness matrix of 
Equation (A. 19) is equivalent to the force-displacement relations for a 
uniform beam that are familiar from classical structural analysis. 

A.3 Member end-releases 

When a structure has an internal pin (i.e., no moment transfer from one 
element to the adjacent element), the DOF associated to the rotation must 
have a stiffness value of zero for that DOF. In that way, the element will 
keep the ability to transfer the axial and shear force but not the bending 
moment. That process of assigning a zero value to one term in the stiffness 
matrix will affect the other terms because equilibrium has to be 
maintained. 



ERDC/ITL TR-16-1 


161 


For a beam element, the six equilibrium equations in the local reference 
system can be written as 


fij= k ij u ij (A.20) 

If one end of the member has a hinge, or other type of release that causes 
the corresponding force to be equal to zero, Equation (A.20) requires 
modification. A typical equation is of the following form: 


12 


fn=Yh u i 

j =1 


(A.21) 


If we know a specific value of f n is zero because of a release, the 
corresponding displacement u n can be written as 


n—1 h- 12 b 

u = V^U.+ V u +r 

n Z_^ 7 j 1 7 1 'n 

j= 1 ^ nn j=n+ 1 ^ nn 


(A.22) 


Therefore, by substitution of Equation (A.22) into the other five 
equilibrium equations, the unknown u n can be eliminated and the 
corresponding row and column set to zero, or 


fij = ki j u ij +n j 


The terms f n = 0 and the new stiffness terms are equal to 


kij 


h 



(A.23) 


(A.24) 


This procedure can be repeatedly applied to the element equilibrium 
equations for all releases. The repeated application of the simple 
numerical equation is sometimes called the static condensation or 
partial Gauss elimination. 

There is a special case when the load is applied at the end release node. In 
this case, the load must be altered to maintain the o moment transfer. This 
special case is discussed in Appendix H. 



ERDC/ITL TR-16-1 


162 


A.4 Element mass matrix 

The mass influence coefficient for a structure is the force in the i th DOF 
due to unit acceleration in the j th DOF. Applying this definition to a beam 
element with distributed mass m (x) and using the principle of virtual 
displacement, a general equation for can be derived: 

L 

m ij= J m { x )Wi{x)Wj{x)dx (A.25) 

o 


The symmetric form of this equation shows that the mass matrix is 
symmetric; . If we use the same interpolation functions of 

Equations (A. 11 ) to (A. 16 ) as were used to derive the element stiffness 
matrix into Equation (A. 20 ), the result obtained is known as the consistent 
mass matrix. The integrals of Equation (A. 20 ) are evaluated numerically 
or analytically depending on the function m(x). For an element with 
uniform mass per unit length (i.e., m(x) = m), the integrals can be 
evaluated analytically to obtain the element (consistent) mass matrix as 


1 

3 

0 


0 


m e = mL 


1 

6 


0 


0 


0 

0 

1 

6 

0 

0 

156 

22 L 

0 

54 

— 13L 

420 

420 

420 

420 

22 L 

4L 2 

0 

13L 

00 

1 

420 

420 

420 

420 

0 

0 

1 

3 

0 

0 

54 

13L 

0 

156 

-22 L 

420 

420 

420 

420 

— 13L 

00 

1 

0 

-22 L 

4L 2 

420 

420 

420 

420 . 


(A.26) 


A.5 Element (applied) force vector 

If the external forces p*(t), i = 1,2, 3,4,5, and 6 are applied along the six 
DOF at the two nodes of the finite element, the element force vector can be 
written directly. On the other hand, if the external forces are concentrated 
forces p 'j(t) at locations Xj, the nodal force in the i th DOF is 



ERDC/ITL TR-16-1 


163 


p,M = X»,<V < a - 27 > 

j 


This equation can be obtained by the principle of virtual displacement. If 
the same interpolation functions of Equations (A. 11 ) to (A. 16 ), are used to 
derive the element stiffness matrix as used here, the results obtained are 
called consistent nodal forces. 

A.6 Nonlinear force-deflection relationship for the springs supports 

The software Impact_Deck has the capability to calculate the response of 
spring supports if the springs develop plastic behavior in the force- 
displacement relationship. The spring can be considered as linear if the 
load in the spring is below the elastic displacement Seias and the elastic 
force Feias as shown in Figure A. 3 . If the load is reduced and the force- 
displacement is below point 1 , the unload returns along the same path as 
the loading phase. The loading phase is shown using green arrows and the 
unloading phase is shown using red arrows. However, if the load is greater 
than the elastic displacement and it is in the loading stage, it follows the 
green arrows until reaching the maximum force-displacement, point 2 . If 
the unload occurs from this point, it will unload following a slope specified 
by the user. In this case, the slope proceeds from point 2 to point 4 . If the 
force never increases to point 2 again, the force-displacement will remain 
along the line from point 2 to point 4 . If the force decreases to point 4 , zero 
force is reached with a plastic permanent deflection. If the load increases 
again until point 2 is reached, the original backbone is rejoined, 
proceeding from point 2 towards point 3 . If the force reaches a maximum 
on the line between point 2 and 3 and starts to decrease again, the load- 
deflection will follow the same unload slope as the slope from point 2 to 
point 4 but starting from the new maximum force-deflection. If the force- 
deflection is greater than point 3 , Impact_Deck assigns a zero value to this 
spring because the maximum value was reached and failure occurs. 



ERDC/ITL TR-16-1 


164 


Figure A.3 Force-displacement relation of the spring 
support. 









ERDC/ITL TR-16-1 


165 


Appendix B: McAlpine Alternative Flexible 
Wall - Equations of Motion for the 
Mathematical Model 

In structural dynamics the mathematical model of bodies of finite 
dimensions undergoing translatory motion are governed by Newton’s 
Second Law of Motion, expressed as 

^F = m«a (1.1 bis) 

at each time step t during motion. In the mathematical model of transverse 
vibration, the forces acting on the flexible impact beam mass at each time 
step t include (l) the impact force at time step t, ( 2 ) the elastic restoring 
forces (of the beam), and ( 3 ) the damping forces (of the beam). The 
mathematical model of the beam in the engineering formulation described 
in this section has a finite number of degrees of freedom (DOF) because it is 
discretized using the finite element formulation. The engineering 
formulation of equations of motion are solved using a numerical solution 
method to determine the displacement and response forces at each pile bent 
support feature. Each group of clustered piles is modeled as transverse, 
longitudinal, and rotational elastic/plastic springs. This is a dynamic 
process, with the response forces responding to an impact pulse force time 
history and each calculation of the equations of motion occurring at each 
time step. The applied impact force can be given a constant velocity parallel 
to the approach wall so that it is changing its position with time. The 
responses of the elastic beam over elastic or plastic springs can be obtained 
by the use of the multiple degrees of freedom model (MDOF) and the finite 
element formulation of the beam element. Impact_Deck includes the 
damping forces by using the Rayleigh damping model. The response is 
obtained using the HHT- a method (Appendix F) or the well-known 
Wilson-0 method. By using the Newton’s Second law and applying it to a 
MDOF system, the resulting equations of motion can be expressed as 

[M]{u(t)J + [C]{u (f)} + [q{u(t)} = {F(0} (2.1 bis) 


where: 


[M] = global mass matrix 



ERDC/ITL TR-16-1 


166 


[C] = global damping matrix 
[K] = global stiffness matrix 
{F(t)} = global vector of external forces and moments 
(u(t)}, {it(t)} = relative acceleration, relative velocity, and relative 
displacement for each DOF. 

This appendix discusses the various relationships used in this engineering 
methodology that are implemented in Impact_Deck. 

B.l Element degrees of freedom (DOF) and interpolation functions 

Consider a straight-beam element of length L, mass per unit length m(x), 
and flexural rigidity El (x). The two nodes by which the 2 -D finite element 
can be assembled into a structure are located at its ends. If only planar 
displacements are considered, each node has three DOF: the longitudinal 
displacement, the transverse displacement, and rotation. 

The longitudinal displacement (i.e., axial direction) of the beam element is 
related to its two DOF: 


(B- 1 ) 

1=1 

where the function defines the displacement of the element due to 

unit displacement u t while constraining other DOF to zero. Thus, 0;(x) 
satisfies the following boundary conditions and are shown in Figure B.l, 

i=l:\V 1 (0) = l,\\) 1 (L) = 0 (B.2) 

/ = 2 : \p 2 (0) = 0, ip 2 (L) —1 (B.3) 

Note that these DOF subscript values i = l and 2 correspond to Figure B.l 
DOF subscripts i = l and 4 . 

The transverse displacement and rotation of the beam element is related 
to its four DOF, 



ERDC/ITL TR-16-1 


167 


Figure B.l Shape function for axial displacement 
effect. 



u(x,t) = jyi(t)Wi(x) i = 2 ,3,5,6 

(B.4) 

where the function i/>;(x) defines the displacement of the element due to 
unit displacement Ui, while constraining other DOF to zero. Thus, ify(x) 
satisfies the following boundary conditions, 

i= 1 •' tpfyOJ = l,ip 1 (0) =tp ± (L) =\Pi (L) = 0 

(B.5) 

i= 2 :tp' 2 ro; = i,\p 2 ( 0 )=ip 2 (L) = \p 2 (L) = 0 

(B. 6 ) 

i= 3 : tp 3 (L) = (0) = W 3 ( 0 ) = W 3 {L) = 0 

(B.7) 

i= 4 ■’W , 4( l ) = 1 ’W4( 0 )=W , 4(0)=W 4 ( l ) = 0 

(B. 8 ) 


Note that these DOF subscript values i = l, 2 , 3 , and 4 correspond to 
Figure B .2 DOF subscripts i = 2 , 3 , 5 , and 6 . 

These interpolation functions could be any arbitrary shapes satisfying the 
boundary conditions. One possibility is the exact deflected shapes of the 
beam element due to the imposed boundary conditions, but these are 
difficult to determine if the flexural rigidity varies over the length of the 
element. However, they can conveniently be obtained for a uniform beam 












ERDC/ITL TR-16-1 


168 


as illustrated next for the transverse displacement and rotation. Neglecting 
shear deformations, the equilibrium equation for a beam loaded only at its 
ends is 


El 


d 4 u 

dx 4 


0 


(B.9) 


The general solution of Equation (B. 9 ) for a uniform beam is a cubic 
polynomial 


u(x) =a 1 +a 2 


V 

+ flo 

/ \ 

X 

2 

+ a a 

X 


0 

L, 


X, 


(B.10) 


The constants a t can be determined for each of the four sets of boundary 
conditions of Equations (B. 5 ) to (B. 8 ), to obtain 


xpi(x) = l - 3 

W 2 ( x ) = l 
W 3 (*) = 3 
T 4 {x) = -L 





( \ 

V 


( \ 

X 

2 

( \ 

X 

L 

A 

-2 L 

+ L 


J, 


L, 

L, 


( \ 

X 

2 

_ 9 

/ \ 

X 

L, 

— z 

L, 


'x 


yL, 


+ L 


'x ' 3 


(B.ll) 


(B.12) 


(B.13) 


(B.14) 


These interpolation functions, illustrated in Figure B. 2 , can be used in 
formulating the element matrices. The same process can be done with the 
axial effect to obtain the basic interpolation functions, 




t 2 (*) = - 


(B.15) 

(B.16) 


The finite element method is based on assumed relationships between the 
displacements at interior points of the element and the displacements at 
the nodes. Proceeding in this manner makes the problem tractable but 
introduces approximations in the solution. 



ERDC/ITL TR-16-1 


169 


Figure B.2 Shape function for transverse 
displacement and rotation effect. 



B.2 Element stiffness matrix 

Consider a beam element of length L with flexural rigidity EI(x). By 
definition, the stiffness influence coefficient k t j of the beam element is the 

force in the DOF i due to unit displacement in DOFj. Using the principle 
of virtual displacement, the general equation for k^, which is the stiffness 

term for the transverse displacement and rotation, in the element stiffness 
matrix is: 


L 

k, = /EI( x)ip ](x)Wj(x)dx (B.17) 

0 

The symmetric form of this equation shows that the element stiffness 
matrix is symmetric; /q ; - = kj t . Equation (B. 17 ) is a general result in the 

sense that it is applicable to elements with arbitrary variation of flexural 
rigidity EI(x), although the interpolation functions of Equations (B. 11 ) to 
(B. 14 ) are exact only for uniform elements. The associated errors can be 
reduced to any desired degree by reducing the element size and increasing 
the number of finite elements in the structural idealization. For a uniform 
finite element with EI(x) = El, the integral of Equation (B. 17 ) can be 

















ERDC/ITL TR-16-1 


170 


B.3 


evaluated analytically for i,j = 2,3,5, and 6, resulting in the 
corresponding terms in the element stiffness matrix. 

In the same way, for the axial contribution, the general equation for k t j, 
which is the stiffness term for the axial displacement in the element 
stiffness matrix is 


L 

ky = fEA(x)\p](x)\p'j(x)dx (B. 18 ) 

0 


For a uniform finite element with EA(x) = EA, the integral of Equation 
(B.18) can be evaluated analytically for i,j = 1,4 resulting in the 
corresponding terms in the element stiffness matrix. Finally, when all 
terms are calculated, the element stiffness matrix can be obtained as 



AE 

0 

0 

AE 

0 

0 


L 

L 


0 

12 EI 

6EI 

0 

12 EI 

6 El 


L 3 

L 2 

L 3 

L 2 

*1= 

0 

AE 

6 El 

L 2 

0 

4 El 

L 

0 

0 

AE 

6EI 

L 2 

0 

2 EI 

L 

0 


L 

L 


0 

12 EI 

6EI 

0 

12 EI 

6 El 


L 3 

L 2 

L 3 

L 2 


0 

6EI 

2 EI 

0 

6EI 

4 EI 


L 2 

L 

L 2 

L 


(B. 19 ) 


These stiffness coefficients are the exact values for a uniform beam, 
neglecting shear deformation, because the interpolation functions are the 
true deflection shapes for this case. Observe that the stiffness matrix of 
Equation (B.19) is equivalent to the force-displacement relations for a 
uniform beam that are familiar from classical structural analysis. 

Member end-releases 

When a structure has an internal pin (i.e., no moment transfer from one 
element to the adjacent element), the DOF associated to the rotation must 
have a stiffness value of zero for that DOF. In that way the element will 
keep the ability to transfer the axial and shear force but not the bending 



ERDC/ITL TR-16-1 


171 


moment. That process of assigning a zero value to one term in the stiffness 
matrix will affect the others terms because equilibrium has to be 
maintained. 

For a beam element, the six equilibrium equations in the local reference 
system can be written as 


/#=V 9 (B-20) 

If one end of the member has a hinge, or other type of release that causes 
the corresponding force to be equal to zero, Equation (B.20) requires 
modification. A typical equation is of the following form: 


12 


j =1 


(B. 21 ) 


If we know a specific value of f n is zero because of a release, the 
corresponding displacement u n can be written as 


7i—i l- 12 k 

u =Y—u-+ T -^H.+r 

n Z_-/ 7 1 Z_-/ Is J n 

j =1 % 


h 

j=n +1 ^nn 


(B. 22 ) 


Therefore, by substitution of Equation (B.22) into the other five 
equilibrium equations, the unknown u n can be eliminated and the 
corresponding row and column set to zero, or 


f ij = k ij U ij + r v 


The terms f n = 0 and the new stiffness terms are equal to 


kij 


k v 



(B. 23 ) 


(B. 24 ) 


This procedure can be repeatedly applied to the element equilibrium 
equations for all releases. The repeated application of the simple 
numerical equation is sometimes called the static condensation or 
partial Gauss elimination. 



ERDC/ITL TR-16-1 


There is a special case when the load is applied at the end release node. In 
this case, the load must be altered to maintain the o moment transfer. This 
special case is discussed in Appendix H. 

B.4 Element mass matrix 

The mass influence coefficient m t j for a structure is the force in the i th DOF 
due to unit acceleration in thej f,! DOF. Applying this definition to a beam 
element with distributed mass m (x) and using the principle of virtual 
displacement, a general equation for can be derived 

L 

m ij= f m (x)Vi(x)Wj(x)dx (B.25) 

o 


The symmetric form of this equation shows that the mass matrix is 
symmetric; = m ;i . If we use the same interpolation functions of 

Equations (B.n) to (B. 16 ) as were used to derive the element stiffness 
matrix into Equation (B. 23 ), the result obtained is known as the consistent 
mass matrix. The integrals of Equation (B. 23 ) are evaluated numerically 
or analytically depending on the function m(x). For an element with 
uniform mass per unit length (i.e., m(x) = m), the integrals can be 
evaluated analytically to obtain the element (consistent) mass matrix as 


m e = mL 


1 

3 

0 

0 

1 

6 

0 

0 

0 

156 

22 L 

0 

54 

— 13L 

420 

420 

420 

420 

0 

22 L 

4L 2 

0 

13L 

-3L 2 

420 

420 

420 

420 

1 

6 

0 

0 

1 

3 

0 

0 

0 

54 

13L 

0 

156 

-22 L 

420 

420 

420 

420 

0 

— 13L 

—3L 2 

0 

-22 L 

4L 2 

420 

420 

420 

420 . 


(B.26) 


B.5 Element (applied) force vector 

If the external forces p;(t), i = 1,2,3,4 ,5 and 6 are applied along the six 
DOF at the two nodes of the finite element, the element force vector can be 


172 



ERDC/ITL TR-16-1 


173 


written directly. On the other hand, if the external forces are concentrated 
forces p at locations Xj, the nodal force in the i th DOF is 

P l (t) = Yj>'jV,(x,) (B-27) 

j 

This equation can be obtained by the principle of virtual displacement. If 
the same interpolation functions of Equations (B.n) to (B. 16 ) are used to 
derive the element stiffness matrix as used here, the results obtained are 
called consistent nodal forces. 

B.7 Nonlinear force-deflection relationship for the springs supports 

The Impact_Deck software has the capability to calculate the response of 
spring supports if the springs develop plastic behavior in the force- 
displacement relationship. The spring can be considered as linear if the load 
in the spring is below the elastic displacement Seias and the elastic force Feias 
as shown in Figure B. 3 . If the load is reduced and the force-displacement is 
below point 1 , the unload process returns along the same path as the loading 
phase. The loading phase is shown using green arrows and the unloading 
phase is shown using red arrows. However, if the load is greater than the 
elastic displacement and it is in the loading stage, it follows the green 
arrows until reaching the maximum force-displacement, point 2 . If the 
unload occurs from this point, it will unload following a slope specified by 
the user. In this case, the slope proceeds from point 2 to point 4 . If the force 
never increases to point 2 again, the force-displacement will remain along 
the line from point 2 to point 4 . If the force decreases to point 4 , zero force 
is reached with a plastic permanent deflection. If the load increases again 
until point 2 is reached, the original backbone curve is rejoined, proceeding 
from point 2 towards point 3 . If the force reaches a maximum on the line 
between point 2 and 3 and starts to decrease again, the load-deflection will 
follow the same unload slope as the slope from point 2 to point 4 , but 
starting from the new maximum force-deflection. If the force-deflection is 
greater than point 3 , Impact_Deck assigns a zero value to this spring 
because the maximum value was reached and failure occurs. 



ERDC/ITL TR-16-1 


174 


Figure B.3 Force-displacement relation of the spring 
support. 



B.8 Transformation of the stiffness matrix from local to global 
coordinate system 

The Impact_Deck computer program has the option to calculate the 
dynamic response of a flexible wall system such as the McAlpine 
alternative flexible wall system when subjected to an impact load. The 
McAlpine alternative flexible wall consists of a series of elastic beams 
supported over a series of clustered pile groups. The beams transfer the 
axial and shear forces to the pile cap by means of a shear key. Due to the 
fact that the beams are not directly connected between the impact beams, 
that is, they are connected by means of the cap beam shear key, the 
mathematical model has to include a rigid beam element to model the 
distance, stiffness, and mass of the shear key. Assuming that a barge train 
moving at 3 ft/sec impacts the flexible wall at the end of a beam, and with 
a time history duration of 3 sec, the whole impact process will have 9 ft of 
length in contact, which occurs inside the shear key model length (i.e., 
contact solely with the concrete cap to the three-pile group) and not over 
the beams. A typical McAlpine alternative flexible wall is presented in 
Figure B. 4 , and this arrangement of the structural system can be observed. 

Impact_Deck also models the effect of a shear key and pile cap along two 
consecutive beams. Two rigid beam elements are used to model the end¬ 
points of the two incoming beams to the center of the pile cap along the 
longitudinal center-line of the beams. However, the triangular arrangement 








ERDC/ITL TR-16-1 


175 


Figure B.4 Typical McAlpine flexible wall system. 


Cross-section 


28 ' 

12 ' 2 " 8 ' 3 " 6 ' 7 " 


t t 




s 


Barge 

J lmpact 





McAlpine Alternative 
Flexible Wall 


0> C 

L8' 

oo 

-_ 

Barge Impact ab | > ! < [j 


O 




Figure B.5 McAlpine flexible wall mathematical model. 



of piles allows the introduction of a possible rotation of the pile cap. That 
rotation occurs at the center of rigidity of the pile cap with respect to its 
supporting piles. In this case, the center of rigidity is at the center of the 
pile cap, but along the transverse direction (perpendicular to the center- 
line of the beams). To see the calculation of the center of rigidity, please 
refer to Appendix E. The correct position of the equivalent translational 
and rotational spring is at this center of rigidity, which does not coincide 
with the longitudinal global axis of the beams. To connect the springs 
located at the center of rigidity to the flexible beam elements and rigid 
beam element (shear key), an additional rigid beam element is included 


























































































ERDC/ITL TR-16-1 


176 


perpendicular to the longitudinal beam (in the transverse direction) and 
connected to a node at the center of rigidity of the pile cap. This concept 
can be visualized in Figure B. 5 . 

This rigid beam element has the same stiffness matrix in local coordinates 
as a general beam element which stiffness matrix is 


AE 

0 

0 

AE 

0 

0 

L 

L 

0 

12EI 

6 EI 

0 

12EI 

6 EI 

L 3 

L 2 

L 3 

L 2 

0 

6 El 

L 2 

4 El 

L 

0 

6 EI 

L 2 

2EI 

L 

AE 

0 

0 

AE 

0 

0 

L 

L 

0 

12EI 

6 EI 

0 

12EI 

6 El 

L 3 

L 2 

L 3 

L 2 

0 

6 EI 

2EI 

0 

6 EI 

4EI 

L 2 

L 

L 2 

L 


If the local axis of the element does not coincide with the global axis of the 
transformation of axis has to be done to assemble the global stiffness 
matrix of the structure. The transformation from local coordinate system 
to global coordinate system, as shown in Figure B .5 can be done by the 
transformation matrix. 

First, let the equilibrium equations in local coordinates be, 

{f'} = [k']{A} (B.29) 

If each side of Equation (B. 29 ) is pre-multipled by transformation matrix, 
it results in, 


[ r ]{F} = [^][r]{A} 


(B.30) 


If Equation (B. 30 ) is again pre-multipled by the transpose of the 
transformation matrix, 


[ r f[r]{F} = [r] r [r][r]{A} 


(B.31) 



ERDC/ITL TR-16-1 


177 


And noting that the transformation matrix is orthonormal, that is, 


[rf [r]=[z] 

(B.32) 

I]{F} = {F} 

(B.33) 


then, 

{F} = [ r f[X'][r]{A} (B.34) 

Equation (B. 34 ) relates the forces in global coordinates to the stiffness 
matrix in local coordinates and the displacements in global coordinates. 
Equation (B. 34 ) can be expressed as 

{F} = \K\{ A} (B.35) 

where [K] is the global stiffness matrix of the element. Finally, the stiffness 
matrix in global coordinates can be expressed as multiplication of three 
matrices, the transpose of the transformation matrix by the stiffness 
matrix in local coordinates, and by the transformation matrix as 

[AT] = [rf [k'\[T] (B.36) 

It can be demonstrated by the equilibrium equations that the 
transformation matrix has the form 

cos0 sinQ 0 0 0 0 

-sm0 cos0 0 0 0 0 

0 0 1 0 0 0 

0 0 0 cos0 sinQ 0 

0 0 0 -sm0 cos0 0 

0 0 0 0 0 1 

The Impact_Deck computer program performs this matrix calculation to 
transform the stiffness matrix in local coordinate system to global 
coordinate system for the rigid beam element which is perpendicular to 
the beam alignment as shown in Figure B. 6 . In that case, Equation (B. 8 ) 
uses a value of 9 = 90 degrees (the angle between the local axis to the 



(B.37) 



ERDC/ITL TR-16-1 


178 


global axis). The same transformation procedure is done for the element 
mass matrix to transform from local to global coordinate system for the 
perpendicular rigid beam element. 


Figure B.6 Transformation of beam element coordinate 
system (Local-Global). 







ERDC/ITL TR-16-1 


179 


Appendix C: Guard wall - Equations of Motion 
for the Mathematical Model 

In structural dynamics the mathematical model of bodies of finite 
dimensions undergoing translatory motion are governed by Newton’s 
Second Law of Motion, expressed as 

= m*a (1.1 bis) 


at each time step t during motion. In the mathematical model of 
transverse vibration, the forces acting on the flexible impact beam mass at 
each time step t include (l) the impact force at time step t, ( 2 ) the elastic 
restoring forces (of the beam), and ( 3 ) the damping forces (of the beam). 
The mathematical model of the beam in the engineering formulation 
described in this has a finite number of degrees of freedom (DOF) because 
it is discretized using the finite element formulation. The engineering 
formulations of equations of motion are solved using a numerical solution 
method to determine the displacement and response forces at each pile 
bent support feature. Each group of clustered piles is modeled as 
transverse and longitudinal elastic/plastic springs. This is a dynamic 
process, with the response forces responding to an impact pulse force time 
history and each calculation of the equations of motion occurring at each 
time step. The applied impact force can be given a constant velocity 
parallel to the approach wall so that it is changing position with time. The 
responses of the elastic beam over elastic or plastic springs can be 
obtained by the use of the multiple degrees of freedom model (MDOF) and 
the finite element formulation of the beam element. Impact_Deck includes 
the damping forces by using the Rayleigh damping model. The response is 
obtained using either the HHT-a method (Appendix F) or the well-known 
Wilson-0 method (Appendix G). By using the Newton’s Second law and 
applying it to a MDOF system, the resulting equations of motion can be 
expressed as 


mwoj+iciwoj+mmom^o} (2.1 bis) 


where: 

[M] = global mass matrix 
[C] = global damping matrix 



ERDC/ITL TR-16-1 


180 


[K] = global stiffness matrix 
{F(t)> = global vector of external forces and moments 
{u(t)}, {u(t)}, {u(t)} = relative acceleration, relative velocity, and relative 
displacement for each DOF. 

This appendix discusses the various relationships used in this engineering 
methodology that are implemented in Impact_Deck. 

C.l Element degrees of freedom (DOF) and interpolation functions 

Consider a straight-beam element of length L, mass per unit length m(x), 
and flexural rigidity El (x). The two nodes by which the 2-D finite element 
can be assembled into a structure are located at its ends. If only planar 
displacements are considered, each node has three DOF: the longitudinal 
displacement, the transverse displacement, and rotation. 

The longitudinal displacement (i.e., axial direction) of the beam element is 
related to its two DOFs: 


U ( X > t ) = Yj U i(. t W X ) (C- 1 ) 

1 = 1 

where the function defines the displacement of the element due to 

unit displacement u i: while constraining other DOF to zero. Thus <pi (x) 
satisfies the following boundary conditions and are shown in Figure C.l, 


i= 1 -Ti(0) = l, W 1 {L) = 0 

(C. 2 ) 

i= 2 ;ip 2 (0) = 0, ip 2 (L)=l 

(C. 3 ) 


Note that these DOF subscript values i = 1 and 2 correspond to Figure C.l 
DOF subscripts i = 1 and 4. 

The transverse displacement and rotation of the beam element is related 
to its four DOF, 



ERDC/ITL TR-16-1 


181 


Figure C.l Shape function for axial displacement 
effect. 



u(x,t) = Y^i(t)Wi(x) i = 2,3,5,6 (C.4) 

where the function i/>;(x) defines the displacement of the element due to 
unit displacement u it while constraining other DOF to zero. Thus, ipi(x) 
satisfies the following boundary conditions, 

i= i :tp 1 ('0J = l,\p' 1 (0)=\p 1 (L)=\p' 1 (L) = 0 

i= 2 ty' 2 (0) — l,\p 2 (0) =ip 2 (L) = ip 2 (L) = 0 

i= 3 : rp 3 (L) = l,ip 3 (0) = ip 3 (0) = ip 3 (L) = 0 

i= 4 • ty\(L) = l,ip 4 (°) =ip' 4 (0) =\p 4 (L) = 0 

Note that these DOF subscript values z = l, 2, 3, and 4 correspond to 
Figure C.2 DOF subscripts i = 2, 3, 5, and 6. 

These interpolation functions could be any arbitrary shapes satisfying the 
boundary conditions. One possibility is the exact deflected shapes of the 
beam element due to the imposed boundary conditions, but these are 
difficult to determine if the flexural rigidity varies over the length of the 
element. However, they can conveniently be obtained for a uniform beam 


(C. 5 ) 

(C.6) 

(C. 7 ) 

(C.8) 












ERDC/ITL TR-16-1 


182 


as illustrated next for the transverse displacement and rotation. Neglecting 
shear deformations, the equilibrium equation for a beam loaded only at its 
ends is 


El 



0 


(C. 9 ) 


The general solution of Equation (C.9) for a uniform beam is a cubic 
polynomial 


u 


M 


V 

+ flo 

/ \ 

X 

2 

+ CL a 

' x' 


0 

L, 

H- 

X, 


(C. 10 ) 


The constants a t can be determined for each of the four sets of boundary 
conditions of Equations (C.5) to (C.8), to obtain 


( -vO 

2 

/ \ 

v 

1-3 

A 

+ 2 

A 


J, 

J, 


(C.ll) 


W 2 { x ) = l 



( \ 

v 


( \ 

X 

2 

( \ 

X 

L 

A 

— 2 L 

+ L 




L, 

~L 


(C. 12 ) 


W 3 (*) = 3 


/ \ 

X 

2 

9 

/ \ 

X 

L, 

— z 

L 


(C. 13 ) 








+ L 


'x x3 




(C. 14 ) 


These interpolation functions, illustrated in Figure C.2, can be used in 
formulating the element matrices. The same process can be done with the 
axial effect to obtain the basic interpolation functions, 


Ti(*) = l — 


(C. 15 ) 


t 2 (*) = - 


(C.16) 



ERDC/ITL TR-16-1 


183 


The finite element method is based on assumed relationships between the 
displacements at interior points of the element and the displacements at 
the nodes. Proceeding in this manner makes the problem tractable but 
introduces approximations in the solution. 


Figure C.2 Shape function for transverse 
displacement and rotation effect. 






C.2 Element stiffness matrix 

Consider a beam element of length L with flexural rigidity EI(x). By 
definition, the stiffness influence coefficient k t j of the beam element is the 

force in the DOF i due to unit displacement in DOFj. Using the principle 
of virtual displacement, the general equation for k t j, which is the stiffness 

term for the transverse displacement and rotation, in the element stiffness 
matrix is 


L 



(C. 17 ) 


o 


The symmetric form of this equation shows that the element stiffness 
matrix is symmetric; ky = kji- Equation (C.19) is a general result in the 

sense that it is applicable to elements with arbitrary variation of flexural 




















ERDC/ITL TR-16-1 


184 


rigidity EI(x), although the interpolation functions of Equations (C.n) to 
(C.14) are exact only for uniform elements. The associated errors can be 
reduced to any desired degree by reducing the element size and increasing 
the number of finite elements in the structural idealization. For a uniform 
finite element with EI(x) = El, the integral of Equation (C.17) can be 
evaluated analytically for i,j = 2,3,5, and 6, resulting in the 
corresponding terms in the element stiffness matrix. 

In the same way, for the axial contribution, the general equation for k t j, 
which is the stiffness term for the axial displacement in the element 
stiffness matrix is 


L 

= fEA(x)\v i (x)\v' j (x)dx (C.18) 

0 


For a uniform finite element with EA(x) = EA, the integral of Equation 
(C.18) can be evaluated analytically for i,j = 1,4 resulting in the corres¬ 
ponding terms in the element stiffness matrix. Finally, when all terms are 
calculated, the element stiffness matrix can be obtained as 



AE 

0 

0 

AE 

0 

0 


L 

L 


0 

12 EI 

6EI 

0 

12 EI 

6EI 


L 3 

L 2 

L 3 

L 2 

*1= 

0 

AE 

6EI 

L 2 

0 

4 El 

L 

0 

0 

AE 

6EI 

L 2 

0 

2 EI 

L 

0 


L 

L 


0 

12 EI 

6EI 

0 

12 El 

6 El 


L 3 

L 2 

L 3 

L 2 


0 

6EI 

2 EI 

0 

6EI 

4 EI 


L 2 

L 

L 2 

L 


(C.19) 


These stiffness coefficients are the exact values for a uniform beam, 
neglecting shear deformation, because the interpolation functions are the 
true deflection shapes for this case. Observe that the stiffness matrix of 
Equation (C.19) is equivalent to the force-displacement relations for a 
uniform beam that are familiar from classical structural analysis. 



ERDC/ITL TR-16-1 


185 


C.3 Member end-releases 

When a structure has an internal pin (i.e., no moment transfer from one 
element to the adjacent element), the DOF associated to the rotation must 
have a stiffness value of zero for that DOF. In that way the element will 
keep the ability to transfer the axial and shear force but not the bending 
moment. That process of assigning a zero value to one term in the stiffness 
matrix will affect the others terms because equilibrium has to be 
maintained. 

For a beam element, the six equilibrium equations in the local reference 
system can be written as 


4 =^-U.. (C. 20 ) 

If one end of the member has a hinge, or other type of release that causes 
the corresponding force to be equal to zero, Equation (C.20) requires 
modification. A typical equation is of the following form: 


12 


fn=T,KjUj 

j=1 


(C. 21 ) 


If we know a specific value of f n is zero because of a release, the 
corresponding displacement u n can be written as 


n—1 b 12 b 

= y\L u + y -2L 

j=1 ^nn j=n +1 %n 


U J+ r n 


(C. 22 ) 


Therefore, by substitution of Equation (C.21) into the other five 
equilibrium equations, the unknown u n can be eliminated and the 
corresponding row and column set to zero, or 


fij = kij Uy +/’// 


The terms f n = 0 and the new stiffness terms are equal to 


kij 


k.. 



(C. 23 ) 


(C. 24 ) 



ERDC/ITL TR-16-1 


186 


This procedure can be repeatedly applied to the element equilibrium 
equations for all releases. The repeated application of the simple 
numerical equation is sometimes called the static condensation or 
partial Gauss elimination. 

There is a special case when the load is applied at the end release node. In 
this case, the load must be altered to maintain the o moment transfer. This 
special case is discussed in Appendix H. 

C.4 Element mass matrix 

The mass influence coefficient for a structure is the force in the i th DOF 
due to unit acceleration in the j th DOF. Applying this definition to a beam 
element with distributed mass m (x) and using the principle of virtual 
displacement, a general equation for can be derived 

L 

m ij= J m { x )Wi{x)Wj{x)dx (C. 25 ) 

o 

The symmetric form of this equation shows that the mass matrix is 
symmetric; j. If we use the same interpolation functions of 

Equations (C.n) to (C.16) as were used to derive the element stiffness 
matrix into Equation (C.20), the result obtained is known as the consistent 
mass matrix. The integrals of Equation (C.20) are evaluated numerically or 
analytically depending on the function m(x). For an element with uniform 
mass per unit length (i.e., m(x) = m), the integrals can be evaluated 
analytically to obtain the element (consistent) mass matrix as 



ERDC/ITL TR-16-1 


187 


1 

3 

0 


0 


m e — mL 


1 

6 


0 


0 


0 

0 

1 

6 

0 

0 

156 

22 L 

0 

54 

— 13L 

420 

420 

420 

420 

22 L 

4L 2 

0 

13L 

00 

1 

420 

420 

420 

420 

0 

0 

1 

3 

0 

0 

54 

13L 

0 

156 

-22 L 

420 

420 

420 

420 

— 13L 

00 

1 

0 

-22 L 

4L 2 

420 

420 

420 

420 . 


(C.26) 


C.5 Element (applied) force vector 

If the external forces p;(t), i = 1 , 2 , 3 , 4 , 5 , and 6 are applied along the six 
DOF at the two nodes of the finite element, the element force vector can be 
written directly. On the other hand, if the external forces are concentrated 
forces at locations Xj, the nodal force in the i th DOF 

Pi( t ) = YfjVi( x j) (C ‘ 27) 

j 

This equation can be obtained by the principle of virtual displacement. If 
the same interpolation functions of Equations (C.n) to (C.i6) are used to 
derive the element stiffness matrix as used here, the results obtained are 
called consistent nodal forces. 

C.6 Nonlinear force-deflection relationship for the springs supports 

Impact_Deck has the capability to calculate the response of spring 
supports if the springs develop plastic behavior in the force-displacement 
relationship. The spring can be considered as linear if the load in the 
spring is below the elastic displacement Seias and the elastic force Feias as 
shown in Figure C.3. If the load is reduced and the force-displacement is 
below point 1, the unload returns along the same path as the loading 
phase. The loading phase is shown using green arrows and the unloading 
phase is shown using red arrows. However, if the load is greater than the 
elastic displacement and it is in the loading stage, it follows the green 
arrows until reaching the maximum force-displacement, point 2. If the 



ERDC/ITL TR-16-1 


188 


unload occurs from this point, it will unload following a slope specified by 
the user. In this case, the slope proceeds from point 2 to point 4. If the 
force never increases to point 2 again, the force-displacement will remain 
along line from point 2 to point 4 until zero force is reached with a plastic 
permanent deflection. If the load increases again until point 2 is reached, 
the original backbone curve is rejoined proceeding from point 2 towards 
point 3. If the force reaches a maximum on the line between point 2 and 3 
and starts to decrease again, the load-deflection will follow the same 
unload slope as the slope from point 2 to point 4, but starting from the 
new maximum force-deflection. If the force-deflection is greater than 
point 3, Impact_Deck assigns a zero value to this spring because the 
maximum value was reached and failure occurs. 


Figure C.3 Force-displacement relation of the spring 
support. 









ERDC/ITL TR-16-1 


189 


Appendix D Push-over analysis for batter-pile 
bent system 

D.l Bent geometry and analytical model used in the push-over 
analysis 

The pipe pile bent of Figure D.l is examined by push-over analysis 
(Ebeling et al. 2012) to determine its load-displacement characteristics. 
The Saul (1968), CPGA analytical model used in the analysis is shown in 
Figure D.2. 1 The pipe pile bent is comprised of 24-inch-diameter, 
concrete-filled pipe piles. Load-displacement plots will be determined by 
push-over analysis for pinned-head and fixed-head conditions. 


Figure D.l Pipe pile approach wall. 



1 Pile numbers are reported in Figure D.2. 
















































































ERDC/ITL TR-16-1 


190 


Figure D.2 CPGA analytical model. 


8.0 



D.2 Pipe pile properties 

Pipe pile properties are presented below: 

Diameter D p of concrete-filled pipe pile = 2.0 ft = 24 in. 

Area Ap of concrete-filled pipe pile = 3.142 ft 2 = 452 in. 2 

Moment of inertia I P of concrete-filled pipe pile = 0.785 ft4 = 16300 in.4 

Radius of gyration = 0.50 ft = 6 in. 

Distance from neutral axis to extreme fiber c = 1.0 ft. = 12 in. 

Modulus of elasticity E c = 504000 ksf = 3500 ksi 

A simple interaction (axial load - moment) diagram is developed to help 
in assessing the conditions where piles reach their moment or axial load 
limits. The interaction diagram is based on the ultimate capacity of the pile 
members. The procedures described in Rangan and Joyce (1992) can be 
used to develop a simple interaction diagram for a concrete-filled pipe pile. 
The interaction diagram points are 






ERDC/ITL TR-16-1 


191 


• Pure axial compression 

• Balance condition (axial compression and bending) 

• Pure bending 

• Pure axial tension 

The value for pure axial compression is based only on the compressive 
strength of the concrete. The 0.375-inch thick steel pipe casing was not 
included in this calculation. The value for pure axial tension is based only 
on the tensile strength of the steel pipe. Balance point and pure moment 
conditions assumes the contribution only of the concrete in compression 
on the compressive side of the neutral axis and contribution only of the 
steel in tension on the tensile side of the neutral axis. 

The interaction diagram for a 24-inch-diameter, concrete-filled pipe pile is 
presented in Figure D.3. 

Figure D.3 Simple interaction diagram for 24-inch-diameter pipe pile. 







ERDC/ITL TR-16-1 


192 


The interaction diagram assumes that the pipe piles in axial compression 
fail as a result of the materials (i.e., concrete and steel) reaching their 
ultimate capacities, rather than by buckling. However buckling 
computations will be needed to assure that this is the case. If buckling 
loads are less than the ultimate axial compressive loads predicted by the 
interaction diagram, then the buckling loads are to be used in the push¬ 
over analysis. 

Piles are generally founded in soils that will not allow them to develop 
their ultimate capacities. It is up to the engineer performing the push-over 
analysis to consider axial load limitations imposed by the foundation 
materials. This example also considers pile axial capacities that are limited 
by side friction and tip resistance provided by the soil foundation. 

D.3 Soil properties 

A sand foundation with a stiffness that varies linearly with depth is 
investigated. The coefficient of subgrade reaction (n/0 is assumed to be 
50 pci. This value for m is corresponds to Terzaghi’s (1955) “recommended” 
value for a moist medium-dense sand and within the scatter considering m 
values cited in other technical literature (e.g., Davisson 1970). There is no 
water table present in this case (i.e., a “dry” site). 

The axial capacity based on soil limitations are 250 kips for piles in tension 
and 1000 kips for piles in compression. 

Relative stiffness factor (T): 


T = 



= 5.4 ft 


D.4 Buckling evaluation 

Buckling loads for the concrete-filled pipe piles are determined using 
methods described in Yang (1996). Figures D.4 and D.5, after Figures 3 
and 9 of Yang (1966), are provided for use in the analysis. 

The coefficient of free standing length m is equal to the free standing 
length L 0 divided by the relative stiffness factor T, or m = L 0 + T = 24 -r 5.4 
= 4 - 45 ■ 



ERDC/ITL TR-16-1 


193 


Figure D.4 (After Figure 3 Yang 1966) Coefficient of critical 
buckling strength. 



Figure D.5 (After Figure 9 Yang 1966) Coefficient decrement of 
buckling strength. 

































































ERDC/ITL TR-16-1 


194 


The critical buckling load assuming no translation can be determined 
using Figure D.4 (after Figure 3 in Yang 1966). 

For pinned-top non-translating pile cap for Figure D.4, the coefficient of 
critical buckling strength [G] is equal to 0.026 and the Euler critical 
buckling load is 


n El 

PcR = ^t JL {G} = 3847 kips 

For fixed-top non-translating pile cap from Figure D.4, the coefficient of 
critical buckling strength [G] is equal to 0.056 and the Euler critical 
buckling load is 


■ CR 


n El r . 

—p-[G] = 7509 kips 


The critical buckling load with translation can be determined using 
Figure D.5 (after Figure 9 of Yang 1966). Entering Figure D.5 with a 
coefficient of free standing length (m) equal to 4.45, the coefficient of 
translation [Gr] is approximately equal to 0.21 for both pinned-head and 
fixed-head piles. The critical buckling load assuming a translation ( Pcra ) is 


p = p 

1 CRA 1 CR 


- 1 

1 

r C ^ 
2 

A 

L 

\r ) 



It is desirable for the push-over analysis to have the critical buckling load 
for various lateral displacements when performing a push-over analysis. 
This has been accomplished with the aid of MathCAD (1998) for pinned- 
head piles (Table D.i) and fixed-head piles (Table D.2). 

D.5 Push-over analysis for pinned-head condition - dry site 

The first push-over analysis is performed for a pinned-head condition at a 
dry site using Saul’s (1968) method and the CPGA software. The results 
are summarized in this section. The first incremental analysis was run 
using an axial stiffness modifier ( C33 ) for the embedded portion of the pile 
with a value of 1.00 for compression piles and 0.50 for tension piles 
where: 



ERDC/ITL TR-16-1 


195 



A 

~S 


Where: 


A = 


PL 

AE 


This assumes that the pile is supported at its tip with all axial load (P) 
transferred to the tip and: 

8 = actual displacement of the pile under axial load (P) 


Table D.l Euler critical buckling load - translating 
pile top - pinned head condition 





ERDC/ITL TR-16-1 


196 


Table D.2 Euler critical buckling load - translating pile top 
- fixed head condition. 


Euler critical buckling load - Translating pile top 

G t :=0.21 Figure 9, Reference 1 
Fixed top 



r = 0.5 °ft 


c := 1.0-ft 


A := 3-in, 4-in.. 14-in 


crA 


(A):=P cr . l-G " A j 


A ^A ) :=A 


P crA( A ) = 

°kip 

< 

< 

5.932-103 

3 

5.407-10 3 


4 

4.881-10 3 


5 

4.355-103 


6 

3.83-103 


7 

3.304-103 


8 

2.778-10 3 


9 

2.253-10 3 


10 

1.727-10 3 


11 

1.202-10 3 


12 

675.846 


13 

150.188 


14 


This is a crude approximation of axial stiffness used to illustrate the push¬ 
over method. The actual value of the stiffness modifier (C33) should be 
determined by appropriate analytical t-z models and/or pile load tests. To 
obtain the maximum moment below the mudline it is necessary to include 
a PMAXMOM data line where: 


PMAXMOM = 


T 



And [H P ], the coefficient of horizontal load for pinned-head conditions, is 
obtained from Figure D.6 (after Figure 7 of Yang 1966) and is equal to 


0.20. 




ERDC/ITL TR-16-1 


197 


Figure D.6, (After Figure 7 Yang 1966) Coefficient of horizontal load 

capacity. 



Therefore: 


PMAXMOM - 



5 4 

— = 21 ft = 324 in. 
0.20 


The “ALLOW” and UNSP” data lines for CPGA do not represent actual 
allowable loads and buckling loads but were included only to obtain pile 
force and displacement results. 

For the first increment of lateral loading, a trial and error process is used 
to determine the lateral load driving Pile #3 to its axial tensile capacity of 
250 kips. This is accomplished with a lateral load of 88 kips. The CPGA 
input and output for this loading increment is presented below. 

































ERDC/ITL TR-16-1 


18 


CPGA INPUT FOR RUN #1 Pinned-head piles 

10 BATTER PILE BENT PINNED TOP FILE:BP4 

15 PROP 3834. 16300. 16300. 453. 0.5 0.0 1 3 

20 PROP 3834. 16300. 16300. 453. 1.0 0.0 2 

30 SOIL NH .050 L 72. 24. 1 TO 3 

40 PIN 1 TO 3 

50 ALLOW R 1000. 242. 1485. 933. 8544. 8544. 1 TO 3 

70 UNSP S 0.6 0.6 500. 500. N 1 TO 3 

80 PMAXMOM 324. 324. 1 TO 3 


CPGA LOAD-DISPLACEMENT OUTPUT FOR RUN #1 Pinned-head piles 

PILE CAP DISPLACEMENTS 


LOAD 

CASE DX DZ R 

IN IN RAD 

1 .307 9E + 01 -.3695E-01 .5633E-02 


Horizontal 
displacement 
equals 3.1 inch 


PILE FORCES IN LOCAL GEOMETRY 


Ml & M2 NOT AT PILE HEAD FOR PINNED PILES 

PILE FI F2 F3 Ml M2 


Maximum moments 
below the mudline 


K K K IN-K IN-K 

1 Pile #3 reaches its axial load capacity of 250 kips 


2 








ERDC/ITL TR-16-1 


199 


By trial-and-error push-over investigations, it can be shown that the next 
failure mechanism will occur due to flexural yielding of the piles below the 
mudline. This will be followed by buckling of Pile 2 followed by buckling of 
Pile 1 as load is shifted from Pile 2 to Pile 1. In the next CPGA run, a low 
stiffness modifier ( C 33 = 0.0001) is given to Pile 3 to eliminate its ability to 
attract axial load. This amounts to releasing Pile 3 in its axial direction 
since it has reached 250 kips of axial tensile capacity. 

The CPGA analysis is performed with a final incremental barge impact 
load of 54 kips producing a total axial load in Pile 2 of: 506.4 + 14.2 = 

520.6 kips (compression). This is less than the 1000 kips axial 
compressive capacity due to skin friction and end bearing. The additional 
barge impact load of 54 kips brings the flexural demand on the piles below 
the mudline to their yield capacities. Referring to the Figure D.3 
interaction diagram for a compressive axial load of 500 kips the flexural 
yield capacity of the piling is 700 ft-k (8400 in.-k) and the flexural demand 
on Pile #2 is 2817 in.-k + 5524 in.-k = 8341 in.-k. This occurs at a total 
lateral displacement of3.i + 6.o = 9.iin. Referring to previous buckling 
calculations, an axial compressive load of 558 kip with 12.0 in. of lateral 
displacement will induce buckling. Therefore, when subjected to an 
additional 3 in. of lateral displacement, buckling of Pile #2 is expected to 
be followed by buckling of Pile #3. It should be noted that CPGA does not 
have the capability to introduce below mudline flexural hinges and 
therefore it will be assumed for the purpose of constructing the load- 
deformation curve that the stiffness of the system remains unchanged 
between the points below mudline where flexural hinging develops and 
buckling takes place. 

A load-displacement plot for the pinned-head bent (solid blue curve) at a 
dry site is presented in Figure D.8 that is located at the end of this 
Appendix. This figure summarizes the resulting load displacement curves 
from this and three other CPGA analyses that will be discussed 
subsequently for pinned-head and fixed-head conditions at dry and wet 
sand sites. These other three push-over analyses will be summarized prior 
to discussing the resulting push-over load-displacement curves so that 
comparisons can be made among the results four analyses. 



ERDC/ITL TR-16-1 


200 


CPGA INPUT FOR RUN #2 Pinned-head piles 

10 CANTILEVER BATTER PILE BENT FILE:BP5 

15 PROP 3834. 16300. 16300. 453. 0.0001 0.0 3 

20 PROP 3834. 16300. 16300. 453. 1.0 0.0 1 2 

30 SOIL NH .047 L 72. 24. 1 TO 3 

40 PIN 1 TO 3 

50 ALLOW R 1000. 242. 1485. 933. 8544. 8544. 1 TO 3 

70 UNSP S 0.6 0.6 500. 500. N 1 TO 3 

80 PMAXMOM 324. 324. 1 TO 14 

CPGA LOAD-DISPLACEMENT OUTPUT FOR RUN #2 Pinned-head piles 

PILE CAP DISPLACEMENTS 

LOAD 

CASE DX DZ R 

IN IN RAD 

1 .5986E+01 .9708E-01 .1889E-01 

PILE FORCES IN LOCAL GEOMETRY 

PILE FI F2 F3 Ml M2 

K K K IN-K IN-K 


D.6 Push-over analysis for fixed-head condition - dry site 

This section summarizes a second push-over analysis conducted for a 
fixed-head condition at a dry site using the CPGA software. The first 
incremental analysis was run using an axial stiffness modifier ( C 33 ) for the 





ERDC/ITL TR-16-1 


201 


embedded portion of the pile with a value of l.oo for compression piles 
and 0.50 for tension piles. This is a crude approximation of axial stiffness 
used to illustrate the push-over method. The actual value of the stiffness 
modifier (C33) should be determined by appropriate analytical t-z models 
and/or pile load tests. To obtain the maximum moment below the mudline 
it is necessary to include a FUNSMOM data line where: 


FUNSMOM = 


T Ln + ClT 


Where [H/\, the coefficient of horizontal load for fixed-head conditions, is 
obtained from Figure D.6 (after Figure 7 of Yang 1966) and is equal to 
0.47, and 

Lo = free standing length = 24 ft 
a = coefficient of effective embedment obtained from Figure D.7 
for “fixed top translating” = 1.7 

Therefore, 


FUNSMOM = 


T L n +aT 

UX— 


5.4 24 + 1.7(5.4) 

(X47 + 2 


= 28.1 ft = 337 in. 


As before, the “ALLOW” and UNSP” data lines for CPGA do not represent 
actual allowable loads and buckling loads but were included only to obtain 
pile force and displacement results. 

For the first increment of lateral loading, a trial-and-error process is used 
to determine the lateral load causing the pile to reach their moment 
capacities at the pile to pile cap connection. This is accomplished with a 
lateral load of 180 kips. CPGA input and output for this loading increment 
is presented below. 



ERDC/ITL TR-16-1 


202 


Figure D.7 (After Figure 2 Yang 1966) Effective embedment of pile 

at buckling. 



CPGA INPUT FOR RUN #1 Fixed-head piles 

10 BATTER PILE BENT FIXED TOP FILE:BF4 

15 PROP 3834. 16300. 16300. 453. 0.5 0.0 1 3 

20 PROP 3834. 16300. 16300. 453. 1.0 0.0 2 

30 SOIL NH .050 L 72. 24. 1 2 3 

40 FIX 1 TO 3 

50 ALLOW R 1000. 242. 1485. 933. 8544. 8544. 123 

70 UNSP S 0.6 0.6 1000. 1000. N 1 2 3 

80 FUNSMOM 337. 337. 1 3 




























ERDC/ITL TR-16-1 


203 


CPGA LOAD-DISPLACEMENT OUTPUT FOR RUN #1 

Fixed-head piles 

PILE CAP DISPLACEMENTS 


LOAD 


CASE DX DY DZ RX RY RZ 


IN IN IN RAD RAD RAD 


1 .2309E+01 .0000E+00 -.1013E+00 .0000E+00 . 

3155E-02 .0000E+00 

PILE FORCES IN LOCAL GEOMETRY 


PILE FI F2 F3 Ml M2 

From Figure B-3 

K K K IN-K IN-K 

(interaction diagram) for a 

1 32.1 .0 -122.1 .0 7002.7 

tensile load of 122 kips 

FUNSMOM .0 -3817.5 

the moment capacity = 

2 32.3 .0 411.6 .0 7046.6 

610 ft-k (7330 in-k). 

FUNSMOM .0 -3846.5 

Therefore moment 

3 33.0 .0 -63.2 .0 7188.6 

demand * moment 

FUNSMOM .0 -3940.4 

capacity at pile to pile cap 

connection. 


For the second increment of lateral loading, a trial-and-error process is 
used to determine the lateral load causing pile #3 to reach its tensile load 
capacity (250 k). This occurs with a lateral load increase of 64 kips. The 
pile-to-pile cap connection is changed from fix to pin to capture the 
yielding that occurred in Run #1. 




ERDC/ITL TR-16-1 


204 



CPGA LOAD-DISPLACEMENT OUTPUT FOR RUN #2 Fixed-head piles 

PILE CAP DISPLACEMENTS 

LOAD 

CASE DX DZ R 
IN IN RAD 

1 .2121E + 01 .14 93E-01 .4192E-02 

PILE FORCES IN LOCAL GEOMETRY 
PILE FI F2 F3 Ml M2 
K K K IN-K IN-K 

1 5.9 .0 18.0 .0 -1920.8 

2 6.0 .0 376.7 .0 -1937.50 

3 6.2 .0 -186.0 .0 -2014.8 


By trial-and-error push-over investigations, it can be seen that next failure 
mechanism will occur due to flexural yielding of the piles below the 
mudline. This will be followed by buckling of Pile 2 followed by buckling of 


Pile #3 has a tensile load 
of 63.2 k from Run #1 and 
186.0 k from Run #2 giving 
a total axial tensile load of 
249.2 k « 250 k. tensile 
capacity reached. 





ERDC/ITL TR-16-1 


205 


Pile l as load is shifted from Pile 2 to Pile l. In the next CPGA run, a low 
stiffness modifier ( C 33 = o.oooi) is given to Pile 3 to eliminate its ability to 
attract axial load. This amounts to releasing Pile 3 in its axial direction 
since it has reached 250 kips of axial tensile capacity. 

The CPGA analysis is performed with a final incremental barge impact 
load of 22 kips producing a total axial load in Pile 2 of 411.6 + 376.7 + 5.8 
= 794.1 kip (compression). This is less than the 1000 kip axial compressive 
capacity due to skin friction and end bearing. The additional barge impact 
load of 22 kips brings the flexural demand on the piles below the mudline 
to their yield capacities. Referring to the Figure D.3 interaction diagram 
for a compressive axial load of 800 kips, the flexural yield capacity of the 
piling is 670 ft-k (8040 in.-k) and the flexural demand on Pile #2 is 
3846 in.-k + 1937 in.-k + 2246 in.-k = 8029 in.-k indicating demand is 
approximately equal to capacity. This occurs at a total lateral displacement 
of 2.3+ 2.1 + 2.4 = 6.8 in. Referring to previous buckling calculations, an 
axial compressive load of 800 kips at about 12.5 in. of lateral displacement 
will cause buckling. Therefore when subjected to an additional 6 in. of 
lateral displacement, buckling of Pile #2 is expected to be followed by 
buckling of Pile #3. It should be noted that CPGA does not have the 
capability to introduce below-mudline flexural hinges; therefore, it will be 
assumed for the purpose of constructing the load-deformation curve that 
the stiffness of the system remains unchanged between the points below 
mudline where flexural hinging develops and buckling takes place. 



ERDC/ITL TR-16-1 


CPGA INPUT FOR RUN #3 Fixed-head piles 

10 BATTER PILE BENT FIXED TOP FILE:BF6 
15 PROP 3834. 16300. 16300. 453. 1.0 0.0 1 2 
20 PROP 3834. 16300. 16300. 453. 0.0001 0.0 3 
30 SOIL NH .050 L 72. 24. 1 TO 3 
40 PIN 1 TO 3 

50 ALLOW R 1000. 242. 1485. 933. 8544. 8544. 1 TO 3 

70 UNSP S 0.6 0.6 500. 500. N 1 TO 3 

80 PMAXMOM 324. 324. 1 TO 3 

90 BATTER 4. 2 TO 3 

100 ANGLE 0.123 

110 PILE 1 0. 0. 0. 

120 PILE 2 7. 0. 0. 

130 PILE 3 14. 0. 0. 

140 LOAD 1 22. 0. 200. 0. 0. 0. 

190 TOUT 1234567 
200 FOUT 1234567 
210 PFO 1 TO 3 


CPGA OUTPUT FOR RUN #3 Fixed-head piles 

PILE CAP DISPLACEMENTS 

LOAD 

CASE DX DZ R 
IN IN RAD 

1 .2407E+01 .9850E-01 .9300E-02 

PILE FORCES IN LOCAL GEOMETRY 
PILE FI F2 F3 Ml M2 
K K K IN-K IN-K 

1 6.7 .0 198.0 .0 -2179.2 

2 6.9 .0 5.8 .0 -2245.7 

3 7.4 .0 -.2 .0 -2398.8 



A load-displacement plot for the fixed-head bent (solid green curve) at a 
dry site is presented in Figure D.8. 







ERDC/ITL TR-16-1 


207 


D.7 Submerged Site 

A submerged sand foundation with a stiffness that varies linearly with 
depth is investigated. The coefficient of subgrade reaction ( m ) is assumed 
to be 30 pci. This value for m is corresponds to Terzaghi’s (1955) 
“recommended” value for a submerged medium-dense sand per Table 3.2 
of Section 3 in Ebeling, et al. (2012). This is sometimes referred to as a 
“wet” site in that report. 

The axial capacity based on soil limitations are 250 kips for piles in tension 
and 1000 kips for piles in compression. 

Relative stiffness factor (T): 


T = 5 


E J p 


6.0 


ft 


D.8 Buckling evaluation 

Buckling loads for the concrete-filled pipe piles are determined using 
methods described in Yang (1996). Figures D.4 and D.5, after Figures 3 
and 9 of Yang (1966), are provided for use in the analysis. 

The coefficient of free standing length m is equal to the free standing 
length L 0 divided by the relative stiffness factor T, or m = Lo+ T = 24 + 6.0 
= 4.0. 

Assuming no translation, the critical buckling load can be determined 
using Figure D.4 (after Figure 3 in Yang 1966). 

For pinned-top non-translating pile cap from Figure D.4, the coefficient of 
critical buckling strength [G] is equal to 0.030 and the Euler critical 
buckling load is 


71 E J P 


■ CR 


[G] = 3256 kips 


For fixed-top non-translating pile cap from Figure D.4, the coefficient of 
critical buckling strength [G] is equal to 0.062 and the Euler critical 
buckling load is 



ERDC/ITL TR-16-1 


208 


n 2 E c I r n 

Pcr = -^[G] = 6728 kips 

The critical buckling load with translation can be determined using 
Figure D.5 (after Figure 9 of Yang 1966). Entering Figure D.5 with a 
coefficient of free standing length (m) equal to 4.0, the coefficient of 
translation [Gr] is approximately equal to 0.20 for both pinned-head and 
fixed-head piles. Assuming a translation ( Pcra ), the critical buckling load is 


P = P 

1 CRA 1 CR 


-1 

1 

f c) 



A 

L 

<r J 



It is desirable for the push-over analysis to have the critical buckling load 
for various lateral displacements when performing a push-over analysis. 
This has been accomplished with the aid of MathCAD (1998) for pinned- 
head piles (Table D.3) and fixed-head piles (Table D.4). 



ERDC/ITL TR-16-1 


Table D.3 Euler critical buckling load - translating pile 
top - pinned head condition. 




ERDC/ITL TR-16-1 


210 


Table D.4 Euler critical buckling load - translating pile 
top - fixed head condition. 


Euler critical buckling load - Translating pile top 


Gj :=0.20 Figure 9, Reference 1 
Fixed top 




0.5 


0.5-ft 


c := 1.0 ft 


A : = 3in,4in.. 14 in 


P crA^ A ) :-P cr'( 1_ G T'- 


A X (A):=A 


P crA ( A ) .yp A _ 2 ^ A ) .jn 


5.383-103 


3 

4.934-10 3 


4 

4.486-10 3 


5 

4.037-10 3 


6 

3.588-103 


7 

3.14-10 3 


8 

2.691-103 


9 

2.243-10 3 


10 

1.794-103 


11 

1.346-10 3 


12 

897.115 


13 

448.557 


14 


D.9 Push-over analysis for pinned-head condition - wet site 

This section summarizes a third push-over analysis conducted for a 
pinned-head condition at a wet site using Saul’s (1968) method and the 
CPGA software. The first incremental analysis was run using an axial 
stiffness modifier (C 33 ) for the embedded portion of the pile with a value of 
1.00 for compression piles and 0.50 for tension piles where: 


C 


33 


A 

J 


Where: 




ERDC/ITL TR-16-1 


211 


A = 


PL 

AE 


This assumes that the pile is supported at its tip with all axial load (P) 
transferred to the tip and: 

8 = actual displacement of the pile under axial load (P) 

This is a crude approximation of axial stiffness used to illustrate the push¬ 
over method. The actual value of the stiffness modifier (C33) should be 
determined by appropriate analytical t-z models and/or pile load tests. To 
obtain the maximum moment below the mudline, it is necessary to include 
a PMAXMOM data line where: 


PMAXMOM = 


T 



Where \H P ], the coefficient of horizontal load for pinned-head conditions, 
is obtained from Figure D.6 (after Figure 7 of Yang 1966) and is equal to 
0.21. 

Therefore: 


T 0 

PMAXMOM = T —i = — = 28.57 ft = 343 in. 

Kl °- 21 


The “ALLOW” and UNSP” data lines for CPGA do not represent actual 
allowable loads and buckling loads but were included only to obtain pile 
force and displacement results. 

For the first increment of lateral loading, a trial-and-error process is used 
to determine the lateral load driving Pile #3 to its axial tensile capacity of 
250 kips. This is accomplished with a lateral load of 88 kips. CPGA input 
and output for this loading increment is presented below. 



ERDC/ITL TR-16-1 


CPGA INPUT 

FOR 

RUN #1 Pinned-head piles 

10 BATTER PILE 

BENT PINNED TOP FILE:BP8 

15 PROP 

3834 . 

16300. 16300. 453. 0.5 0.0 1 3 

20 PROP 

3834. 

16300. 16300. 453. 1.0 0.0 2 

30 SOIL 

NH 

.030 

L 72. 24. 1 TO 3 

40 PIN 1 

TO 

3 



50 ALLOW 

R 

1000. 

242. 1485. 933. 8544. 8544. 1 TO 3 

70 UNSP 

S 0 

. 6 

0 . 

6 500. 500. N 1 TO 3 

80 PMAXMOM 

343 


343. 1 TO 3 

90 BATTER 4 

. 2 

TO 3 

100 ANGLE 0 

. 1 

2 

3 

110 PILE 

1 

0 . 

0 . 

0 . 

120 PILE 

2 

7 . 

0 . 

0 . 

130 PILE 

3 

14 . 

0 

. 0 . 

140 LOAD 

1 

88. 

0 

. 200. 0. 0. 0. 

190 TOUT 

1 

2 3 

4 

5 6 7 

200 FOUT 

1 

2 3 

4 

5 6 7 

210 PFO 

1 

TO 

3 





PILE CAP DISPLACEMENTS 

LOAD 

CASE DX DZ R 


Horizontal 
displacement 
equals 3.1 inch 


IN IN. RAD 


1 .30 96E+01 -.37 8 9E-01 .5660E-02 




PILE FORCES IN LOCAL GEOMETRY 


Maximum moments 
below the mudline 


LOAD CASE 


1 






ERDC/ITL TR-16-1 


213 


PILE FI F2 F3 Ml M2 

K K K IN.-K IN.-K 


1 7.9 .0 -45.7 .0 -2700.6 

2 8.0 .0 508.5 .0 -2728.53 

3 8.2 .0 -251.2 .0 -2829.1 

Pile #3 reaches its axial load 
capacity of 250 kips 


~k ~k ~k ~k ~k ~k 


By trial-and-error push-over investigations, it can be shown that the next 
failure mechanism will occur due to flexural yielding of the piles below the 
mudline. This will be followed by buckling of Pile 2 followed by buckling of 
Pile 1 as load is shifted from Pile 2 to Pile 1. In the next CPGA run a low 
stiffness modifier (C 33 = 0.0001) is given to Pile 3 to eliminate its ability to 
attract axial load. This amounts to releasing Pile 3 in its axial direction 
since it has reached 250 kips of axial tensile capacity. 

The CPGA analysis is performed with a final incremental barge impact load 
of 52 kips producing a total axial load in Pile 2 of 508.5 + 13.8 = 522.3 kips 
(compression). This is less than the 1000 kips axial compressive capacity 
due to skin friction and end bearing. The additional barge impact load of 
52 kips brings the flexural demand on the piles below the mudline to their 
yield capacities. Referring to the Figure D.3 interaction diagram for a 
compressive axial load of 500 kips, the flexural yield capacity of the piling is 
700 ft-k (8400 in.-k) and the flexural demand on Pile #2 is 2728.5 in.-k + 
5630.7 in.-k = 8359.2 in.-k. This occurs at a total lateral displacement of 
3.1 + 6.3 = 9.4 in. Referring to previous buckling calculations, an axial 
compressive load of 560 kip with 12.0 in. of lateral displacement will induce 
buckling. Therefore, when subjected to an additional 3 in. of lateral 
displacement, buckling of Pile #2 is expected to be followed by buckling of 
Pile #3. It should be noted that CPGA does not have the capability to 
introduce below mudline flexural hinges; therefore, it will be assumed for 



ERDC/ITL TR-16-1 


214 


the purpose of constructing the load-deformation curve that the stiffness of 
the system remains unchanged between the points below mudline where 
flexural hinging develops and buckling takes place. 


CPGA INPUT 

FOR 

RUN 

#2 Pinned-head piles 

10 

CANTILEVER BATTER 

PILE BENT FILE:BP5 

15 

PROP 3834. 16300. 

16300. 

453. 0.0001 0.0 3 

20 

PROP 3834. 16300. 

16300. 

453. 1.0 0.0 1 2 

30 

SOIL NH 

.030 

L 72. 

24 . 1 

TO 3 

40 

PIN 1 TO 

3 




50 

ALLOW R 

1000. 

242 . 

1485. 

933. 8544. 8544. 1 TO 3 

70 

UNSP S 0 

.6 0. 

6 500 

. 500. 

N 1 TO 3 

80 

PMAXMOM 

343. 

343. 

1 TO 14 

90 

BATTER 4 

. 2 TO 3 



100 

ANGLE 0 

. 1 2 

3 



110 

PILE 1 

0 . 0 . 

0 . 



120 

PILE 2 

7. 0. 

0 . 



130 

PILE 3 

14. 0 

. 0 . 



140 

LOAD 1 

52. 0 

. 200 

. 0 . 0 . 

0 . 

190 

TOUT 1 

2 3 4 

5 6 

7 


200 

FOUT 1 

2 3 4 

5 6 

7 


210 

PFO 1 TO 3 





~k ~k ~k ~k ~k ~k 


PILE CAP DISPLACEMENTS 


LOAD 

CASE DX DZ R 

IN IN. RAD 

1 .62 65E+01 .9716E-01 .1972E-01 








ERDC/ITL TR-16-1 


215 


PILE FORCES IN LOCAL GEOMETRY 

LOAD CASE - 1 

PILE FI F2 F3 Ml M2 

K K K IN.-K IN.-K 

1 15.9 .0 195.3 .0 -5464.0 

2 16.4 .0 13.8 .0 -5630.7 

3 17.4 .0 -.5 .0 -5981.1 




A load-displacement plot for the pinned-head bent (dashed blue curve) at 
the wet site is presented in Figure D.8. 

D.10 Push-over analysis for fixed-head condition - wet site 

This section summarizes a fourth push-over analysis conducted for a 
fixed-head condition at a wet site using Saul’s (1968) method and the 
CPGA software. The first incremental analysis was run using an axial 
stiffness modifier (C33) for the embedded portion of the pile with a value of 
1.00 for compression piles and 0.50 for tension piles. This is a crude 
approximation of axial stiffness used to illustrate the push-over method. 
The actual value of the stiffness modifier (C33) should be determined by 
appropriate analytical t-z models and/or pile load tests. To obtain the 
maximum moment below the mudline, it is necessary to include a 
FUNSMOM data line where: 


FUNSMOM = 


T Ln + ClT 


Where \Hf\, the coefficient of horizontal load for fixed-head conditions, is 
obtained from Figure D.6 (after Figure 7 of Yang 1966) and is equal to 
0.52, and: 


Lo = free standing length = 24 ft 



ERDC/ITL TR-16-1 


216 


a = coefficient of effective embedment obtained from Figure D.y 
for “fixed top translating” = 1.75 


Therefore, 


FUNSMOM = 



Lq + uT 
2 


6.0 24 + 1.7(6.0) 

052 + 2 


= 28.6 ft = 344 in. 


As before, the “ALLOW” and UNSP” data lines for CPGA do not represent 
actual allowable loads and buckling loads but were included only to obtain 
pile force and displacement results. 

For the first increment of lateral loading, a trial-and-error process is used 
to determine the lateral load causing the pile to reach their moment 
capacities at the pile to pile cap connection. This is accomplished with a 
lateral load of 180 kips. CPGA input and output for this loading increment 
is presented below. 


CPGA INPUT 

FOR RUN 

#1 Fixed-head piles 

10 BATTER PILE 

BENT 

FIXED TOP FILE:BF7 

15 PROP 3834. 

16300. 

16300. 453. 0.5 0.0 1 3 

20 PROP 3834. 

16300. 

16300. 453. 1.0 0.0 2 

30 SOIL NH 

.030 L 72 

. 24. 1 2 3 

40 FIX 1 TO 

3 



50 ALLOW R 

1000. 242 

. 1485. 933. 8544. 8544. 123 

70 UNSP S 0 

. 6 

0.6 1000. 1000. N 1 2 3 

80 FUNSMOM 

344 

. 344 . 

1 3 

85 FUNSMOM 

344 

. 344. 

2 

90 BATTER 4 

. 2 

TO 3 


100 ANGLE 0 

. 1 

2 3 


110 PILE 1 

0 . 

0 . 0 . 


120 PILE 2 

7 . 

0 . 0 . 


130 PILE 3 

14 . 

0 . 0 . 


140 LOAD 1 

180 

. 0. 200. 0. 0. 0. 

190 TOUT 1 

2 3 

4 5 6 

7 

200 FOUT 1 

2 3 

4 5 6 

7 

210 PFO 1 

TO 

3 





ERDC/ITL TR-16-1 


217 


***************************************************************** 

* * * * * * 

PILE CAP DISPLACEMENTS 

LOAD 

CASE DX DY DZ 

IN IN IN 

1 .2462E+01 .0000E+00 -.1102E+00 




From Figure B-3 (interaction 
diagram) for a tensile load of 
122 kips the moment capacity « 
610 ft-k (7330 in-k). Therefore 
moment demand « moment 
capacity at pile to pile cap 
connection. 

FUNSMOM .0 -3699.8 

2 31.5 .0 431.9 .0 7090.5 

FUNSMOM .0 -3729.1 

3 32.1 .0 -72.9 .0 7232.5 

FUNSMOM .0 -3820.5 


PILE FORCES IN LOCAL GEOMETRY 

LOAD CASE - 1 

PILE FI F2 F3 Ml M2 

K K K IN.-K IN.-K 

1 31.2 .0 -132.9 .0 7045.0 




'k'k'k'k'k'k'k'k'k'k 


For the second increment of lateral loading, a trial-and-error process is 
used to determine the lateral load causing pile #3 to reach its tensile load 
capacity (250 k). This occurs with a lateral load increase of 70 kips. The 



ERDC/ITL TR-16-1 


218 


pile-to-pile cap connection is changed from fix to pin to capture the 
yielding that occurred in Run #1. 


CPGA INPUT 

FOR 

RUN #2 Fixed-head piles 

10 

BATTER PILE BENT FIXED TOP FILE:BF8 

15 

PROP 3834. 16300. 16300. 453. 0.5 0.0 1 3 

20 

PROP 3834. 16300. 16300. 453. 1.0 0.0 2 

30 

SOIL NH 

.030 

L 72. 24. 1 TO 3 

40 

PIN 1 TO 

3 


50 

ALLOW R 

1000. 

242. 1485. 933. 8544. 8544. 1 TO 3 

70 

UNSP S 0 

.6 0. 

6 500. 500. N 1 TO 3 

80 

PMAXMOM 

344 . 

344. 1 TO 3 

90 

BATTER 4 

. 2 TO 3 

100 

ANGLE 0 

. 1 2 

3 

110 

PILE 1 

0 . 0 . 

0 . 

120 

PILE 2 

7. 0. 

0 . 

130 

PILE 3 

14 . 0 

. 0 . 

140 

LOAD 1 

70. 0 

. 200. 0. 0. 0. 

190 

TOUT 1 

2 3 4 

5 6 7 

200 

FOUT 1 

2 3 4 

5 6 7 

210 

PFO 1 TO 3 



~k ~k ~k ~k ~k ~k 


PILE CAP DISPLACEMENTS 

LOAD 

CASE DX DZ R 

IN IN. RAD 

1 .2 4 37E+01 -.217 9E-02 .4668E-02 








ERDC/ITL TR-16-1 


219 


PILE FORCES IN LOCAL GEOMETRY 

LOAD CASE - 1 

PILE FI F2 F3 Ml M2 

K K K IN.-K IN.-K 

1 6.2 .0 -2.6 .0 -2131.9 

2 6.3 .0 419.3 .0 -2151.91 

3 6.5 .0 -207.2 .0 -2235.1 


Pile #3 has a tensile load 
of 32.1 k from Run #1 and 

207.2 k from Run #2 giving 
a total axial tensile load of 

239.3 k « 250 k. tensile 
capacity reached. 


~k ~k ~k ~k ~k ~k 


By trial-and-error push-over investigations it can be seen that next failure 
mechanism will occur due to flexural yielding of the piles below the 
mudline. This will be followed by buckling of Pile 2 followed by buckling of 
Pile l as the load is shifted from Pile 2 to Pile 1. In the next CPGA run, a 
low stiffness modifier ( C33 = 0.0001) is given to Pile 3 to eliminate its 
ability to attract axial load. This amounts to releasing Pile 3 in its axial 
direction since it has reached 250 kips of axial tensile capacity. 

The CPGA analysis is performed with a final incremental barge impact 
load of 22 kips producing a total axial load in Pile 2 of 431.9 + 419.3 + 5.9 
= 857.1 kips (compression). This is less than the 1000 kips axial 
compressive capacity due to skin friction and bearing. The additional 
barge impact load of 22 kips brings the flexural demand on the piles below 
the mudline to their yield capacities. Referring to the Figure D.3 
interaction diagram for a compressive axial load of 800 kips the flexural 
yield capacity of the piling is 675 ft-k (8100 in. k) and the flexural demand 
on Pile #2 is 3729 in. k + 2152 in. k + 2385 in. k = 8266 in. k which 
indicates that the demand is slightly greater than the capacity. This occurs 
at a total lateral displacement of 2.5 + 2.4 + 2.6 = 7.5 in. Referring to 
previous buckling calculations an axial compressive load of 857 kips at 
about 13 in. of lateral displacement will cause buckling. Therefore when 
subjected to an additional 5.5 in. of lateral displacement, buckling of Pile 
#2 is expected to be followed by buckling of Pile #3. It should be noted 
that CPGA does not have the capability to introduce below mudline 



ERDC/ITL TR-16-1 


220 


flexural hinges and therefore it will be assumed for the purpose of 
constructing the load-deformation curve that the stiffness of the system 
remains unchanged between the points below mudline where flexural 
hinging develops and buckling takes place. 


CPGA INPUT 

FOR 

RUN #3 Fixed-head piles 

10 BATTER 

PILE 

BENT FIXED TOP FILE:BF6 

15 PROP 3834. 

16300. 16300. 453. 1.0 0.0 1 2 

20 PROP 3834. 

16300. 16300. 453. 0.0001 0.0 3 

30 SOIL NH 

.030 

L 72. 24. 1 TO 3 

40 PIN 1 TO 3 



50 ALLOW R 

1000. 

242. 1485. 933. 8544. 8544. 1 TO 3 

70 UNSP S 

0.6 

0 . 

6 500. 500. N 1 TO 3 

80 PMAXMOM 

344 


344. 1 TO 3 

90 BATTER 

4. 2 

TO 3 

100 ANGLE 

0 . 1 

2 

3 

110 PILE 1 

0 . 

0 . 

0 . 

120 PILE 2 

7 . 

0 . 

0 . 

130 PILE 3 

14 . 

0 

. 0 . 

140 LOAD 1 

22 . 

0 

. 200. 0. 0. 0. 

190 TOUT 1 

2 3 

4 

5 6 7 

200 FOUT 1 

2 3 

4 

5 6 7 

210 PFO 1 

TO 3 




***************************************************************** 

****** 

PILE CAP DISPLACEMENTS 

LOAD 

CASE DX DZ R 

IN IN RAD 

1 .2 645E+01 .984 9E-01 .9010E-02 

***************************************************************** 






ERDC/ITL TR-16-1 


221 


PILE FORCES IN LOCAL GEOMETRY 

LOAD CASE - 1 

PILE FI F2 F3 Ml M2 

K K K IN.-K IN.-K 


1 6.7 .0 198.0 .0 -2314.0 

2 6.9 .0 5.9 .0 -2384.5 

3 7.4 .0 -.2 .0 -2545.1 




A load-displacement plot for the fixed-head bent (dashed green curve) at 
the wet site is presented in Figure D.8. The resulting structural system 
versus displacement plots characterizes the potential energy capacity of 
the particular batter pile bent being analyzed. The push-over results for 
four systems are shown in this figure. 

In Figure D-8 the load-displacement results by the Saul (1968) method for 
a dry site (m = 50 pci) are represented by solid line and those for a 
submerged or wet site (m = 30 pci) by dashed lines. Yang (1966) and 
COM624G methods cannot be used for batter-pile systems because they 
are only applicable to single vertical pile analysis. The methods can be 
used for systems comprised of multiple vertical piles since a single pile 
from the system can be analyzed and the load-displacement results for the 
entire system derived based on the behavior of that single pile. 

The load-displacement curves for the fixed-head pile system have four 
break points designating places where pile or soil yielding occurs. The 
number of yield points and the type of yielding will be pile bent and 
foundation dependent. For the particular pile-bent-foundation system 
investigated, the first break point (one with lowest displacement demand) 
occurs when flexural yielding takes place at the pile-to-pile cap 
connection. The second breakpoint occurs when Pile 3 yields in axial 



ERDC/ITL TR-16-1 


222 


tension (a foundation to pile transfer mechanism). 1 The third breakpoint 
occurs when flexural yielding takes place in the piles below the mudline. 
The fourth breakpoint occurs when Pile 2 buckles. Pile buckling quickly 
results in pile-bent system failure with little reserve potential energy 
capacity in the system. 



There is little difference between the behaviors of submerged (wet) sites 
and dry sites, recognizing of course that lock approach wall bent systems 
will always be submerged. With batter-pile systems, the resistance to 


1 Recall pile numbers are reported in Figure D.2. 





























ERDC/ITL TR-16-1 


223 


lateral load comes principally from pile axial stiffness and not from 
flexural stiffness, as is the case with vertical pile systems. Therefore, 
changes in lateral subgrade resistance (e.g., m) have little effect on system 
load-displacement behavior. 

It can easily be recognized from Figure D-8 that the fixed-head system 
(green curves) has much greater potential energy capacity than the free- 
head system (blue curves). The free-head system does not possess the 
added lateral force resistance provided by rigid pile-to-pile cap 
connections (which is the first break point for the fixed-head system). 

The information contained in this appendix illustrates the push-over 
analysis for a pinned-head and fixed-head batter pile bent using the Corps 
computer program CPGA (X0080) for wet and dry sites (i.e., m = 30 pci 
and 50 pci, respectively). Note that potential failure mechanisms and the 
sequence in which they form will likely be different for other batter-pile 
bent system groups and pile configurations. 



ERDC/ITL TR-16-1 


224 


Appendix E: Formulation for the rotational 
spring stiffness for the McAlpine flexible 
approach wall clustered group of vertical 
piles model 

E.l Calculation of the Center of Rigidity of a pile group for the 
McAlpine flexible wall model 

The McAlpine flexible wall system is supported over pile groups consisting 
of three piles in a triangular arrangement as shown in Figure E.l. 
Impact_Deck computer program has the capability to perform a dynamic 
analysis of this structure that introduces an important change to the finite 
element model from the other two types of approach wall (i.e., guard wall 
and impact deck) structural configurations. This model change is the 
inclusion of a rotational spring at the central rigid link, in addition to 
moving the two translational elastic-plastic springs to the central rigid link. 
This addition is needed because the piles in the group are not in line with 
each other and a torsional resistance is provided by this configuration, 
which means that the pile bent rotates about a Center of Rigidity (C.R.), 
which is not in line with the wall beam. First, the location of the C.R. 
calculation must be performed because the C.R. position identifies the 
length of the rigid element which is perpendicular to the beam alignment. 


Figure E.l. Plan view of the McAlpine flexible alternative approach wall system. 



The length of the rigid element, which is perpendicular to the beam 
elements, is calculated as the distance between C.R. and the central blue 






























ERDC/ITL TR-16-1 


225 


cross as shown in Figure E.2. Figure E.2 also shows the relation between 
the pile cap local axis and the beam global axis. This Figure also 
demonstrates that the alignment of the centerline of the beam does not 
coincide with the location of the piles. 


Figure E.2 Relation between Global-Axis and central support Local-Axis. 



Figure E.3 shows the dimensions and the distances needed to calculate the 
center of rigidity 


Figure E.3 Location of the Center of Rigidity. 





































ERDC/ITL TR-16-1 


226 


The equations to calculate the C.R. are the following. First, the moment of 
inertia for an element with a circular cross-sectional is 


t ^ j4 

I = — * a 

64 


(E.l) 


The translational elastic stiffness of the pile fixed at bottom and top is 


Kile = 


12 * E * / 


(E. 2 ) 


If the piles have the same cross-sectional area, same modulus of elasticity, 
and same length then, following the notation presented in Figure E.3, the 
coordinates of the Center of Rigidity are 




v _ i=l 

^CR r 


E k n 

i=1 
n 

I Z k xi*yi 


T/ - _ i =1 

1 CR — 7 


T, k xi 


i=1 


(E. 3 ) 


Following the same notation, the equivalent translational spring stiffness 
in the global coordinate system are calculated as 

n 

( k x)„=E k x> (E.4) 

7=1 

n 

{ k r)«,=E k r, (E.5) 

1=1 

where n = number of piles . Finally, the length of the rigid element which 
is perpendicular to the beam can be calculated as 

T — X —p 

^Rigid Link CR C A' 


(E. 6 ) 



ERDC/ITL TR-16-1 


227 


E.2 Numerical example for the calculation of the Center of Rigidity 

The definitions of the variables are presented in Figure E.3. 

Data : 

d = 5 ft 8 in. = pile diameter 
f'c = 5,000 psi 

E = 57,000*(5,000) 1/2 = 4,030,508.65 psi = 580,393.25 ksf = Modulus of elasticity 
L = 20 ft = height of the pile above ground 
ex= beam width/2 = 3 ft- 3.5 in. 

Calculations : 

I = n* (5.6666) 4 / 64 = 50.613 ft 4 

kpile = 12* 580,393.25* 50.613/ (20) 3 

= 44,063.16 kip /ft 

kxi = k X 2 = k X 3 = 44,063.16 kip /ft 

kyi = ky2 = ky3 = 44,063.16 kip /ft 

xi = (28-18)/2 + 18 = 23ft 

x 2 = (28-18)/2 = 5ft 

x 3 = (28-18)/2 = 5ft 

yi = 19.5/2 = 9.75 ft 

y 2 = (19.5 -11.5)/2 +11.5 = 15.5 ft 

y 3 = (19.5 -11.5)/2 = 4 ft 


X C r = (44,063.16 *(23 + 5 + 5)) /132,189.5 



ERDC/ITL TR-16-1 


228 


= lift 

Y C r = (44,063.16 *(9.75+15.5+4))/132,189.5 = 9.75ft 

(k x ) eq . = 132,189.5 kip /ft 

(k Y )eq. = 132,189.5 kip/ft 

l-Rigid Link = 11 - 3.291666 = 7.708333 ft 

E.3 Calculation of the rotational spring stiffness of a pile group for 
the McAlpine flexible wall model 

The rotational stiffness is developed in a pile group when the response 
forces are not aligned to the center of rigidity, forming moment arms from 
the line of action of the forces and the center of rigidity. That concept is 
presented in Figure E.4. Each one of the forces contributes to the resultant 
moment around the axis normal to the plan view. 


Figure E.4 Definition of the forces and 
distances generated when the pile cap rotate. 



The resultant moment creates a rotation in the structure. This angle of 
rotation is associated to the ratio of the lateral displacement to the 
distance from the center of each pile to the center of rigidity, as shown in 
Figure E.5. 











ERDC/ITL TR-16-1 


229 


Figure E.5 Rotational angle definition 
when the pile cap rotate. 



The governing equations to calculate the rotational spring stiffness are the 
following. The distance from the line of action of the shear forces in the 
piles to the center of rigidity are, 

d 1 = x 1 -X CR (E.7) 

(E.8) 

The resultant moment (torsion) due to the shear forces in the piles is 


i=l 


(E.9) 


where the forces in pile number one can be defined as the translational 
stiffness times the displacement achieved, 

Fi = kj * Sj ( E .1Q) 


and the rotation of the pile cap can be calculated as 









ERDC/ITL TR-16-1 


230 


this assumes that all piles will rotate the same amount (i.e., rigid body 
motion). Following that concept, the lateral displacement and shear forces 
at piles two and three are 


S2 = 0 * d 2 

(E.12) 

F 2 =k 2 *S 2 

(E.13) 

-X- 

II 

- ^ 

(E.14) 

F 3 =k 3 *S 3 

(E.15) 

Finally, the rotational spring stiffness can be calculated as the ratio of the 
resultant moment divided by the angle of rotation, as 

l c M C r 

K ~ e 

(E.16) 


E.4 Numerical example of the calculation of the rotational spring 
stiffness of a pile group for the McAlpine flexible wall model 

Using the numerical results obtained in section E.2, the calculation of the 
center of rigidity is presented next. 

Data: 


kxi = k X 2 = k X 3 = 44,063.16 kip /ft 


kyi = ky2 = ky3 = 44,063.16 kip /ft 


di = 23 -11 = 12 ft 


d 2 = d 3 = (5.75 2 + 6 2 ) 1/2 = 8.310ft 


B= 0.063 ft 
Calculations: 


Fi = 44,063.16 * 0.063 = 2,775.98 kips 



ERDC/ITL TR-16-1 


231 


tan 0 = 0.063/12 = 0.00525 

0 = arctan (0.00525) = 0.00525 (small angle) 

d 2 = 0.00525 * 8.310 = 0.0436ft 

F 2 = 44,063.16* 0.0436 = 1,921.15 kips 

d 3 = 0.00525 * 8.310 = 0.0436 ft 

F 3 = 44,063.16* 0.0436 = 1,921.15 kips 

Mcr = 2,775.98* 12 + 1,921.15* 8.310 

+ 1,921.15* 8.310 = 65,241.27 kip * ft 

k r = 65,241.27/ 0.00525 = 12,426,909 kip * ft/rad 



ERDC/ITL TR-16-1 


232 


Appendix F: HHT-a method 

In many structural dynamics applications, only low mode response is of 
interest. For these cases, the use of implicit unconditionally stable 
algorithms is generally preferred over conditionally stable algorithms. For 
unconditionally stable algorithms, a time step may be selected 
independent of stability considerations, thus results in a substantial saving 
of computational effort. In addition to being unconditionally stable, when 
only low mode response is of interest, it is often advantageous for an 
algorithm to possess some form of numerical dissipation to damp out any 
spurious participation of the higher modes. Examples of algorithms 
commonly used in structural dynamics which possess these properties are 
the Wilson-0 method and the Newmark-(3 method restricted to parameter 
1 (V+-) 

values of y > - and /? > ^ . The Newmark family of methods allows the 

amount of dissipation to be continuously controlled by a parameter other 
than the time step. On the other hand, the dissipative properties of this 
family of algorithms are considered to be inferior to the Wilson method, 
since the lower modes are affected too strongly (Hilber et al. 1977 ). 

In the Wilson-0 method, 0 must be selected greater than or equal to 1.37 
to maintain unconditional stability. It is recommended to use a value of 
9 = 1.42, as further increasing 0 reduces accuracy and further increases 
dissipation; but even for 9 = 1.42 the method is highly dissipative. A well 
known deficiency of the Wilson-Q method is that it is generally too 
dissipative in the lower modes, requiring a time step to be taken that is 
smaller than that needed for accuracy. In addition, the Wilson-6 method 
tends to damp out the higher modes and could produce large errors when 
contributions of higher modes are significant (Wilson 2010 ). Therefore, 
the use of Wilson-0 method has limited applications. Despite its 
shortcoming, the Wilson-0 method is considered by many to be the best 
available unconditionally stable one-step algorithm when numerical 
dissipation is desired. 

Since it seemed that the commonly used unconditionally stable, dissipative 
algorithms of structural dynamics all possessed some drawbacks, in 1977 
Hilber, Hughes, and Taylor presented a method called the HHT-a method. 
They were looking for an improved one-step method with the following 
requirements: a) it should be unconditionally stable when applied to linear 
problems, b) it should possess numerical dissipation which can be 



ERDC/ITL TR-16-1 


233 


controlled by a parameter other than the time step, (i.e., no numerical 
dissipation should be possible), and c) the numerical dissipation should 
not affect the lower modes too strongly. The resulting new algorithm, 
which consists of a combination of positive Newmark p-dissipation and 
negative a-dissipation, is shown to have improved characteristics when 
compared to the Wilson-0 method. 

The Hilber, Hughes, and Taylor HHT-a method is a generalization of the 
Newmark-P method. The finite-difference equations for the HHT-a 
method are identical to those of the Newmark -(3 method. However, the 
equations of motion has to be modified using the parameter a, as follows, 

mii t + M +1 1 + a)eu t+ M ~ acu > +1 1 + a ) ku t + M ~ aku t = { 1 + a )f t+ ^ (F.l) 

where a, (5, and yare free parameters, which govern the stability and 
numerical dissipation of the algorithm. If a = o this family of algorithms 
reduces to the Newmark family. In this case if y = ^ the algorithms possess 

no numerical dissipation whereas if y > - numerical dissipation is present 

m 2 

and if /? > -—— the new algorithm is unconditionally stable. However, by 

appropriately combining negative a-dissipation with particular values of J3 
and y, a one-parameter family of algorithms with the attributes previously 
enumerated can be developed. Hilber, Hughes, and Taylor in 1977 used /? = 

^ tr') ^ 1 

—-— and y = - — a. They found by numerical experimentation that the 
range of practical interest was — ^ < a < 0. This ensures adequate 

dissipation in the higher modes and at the same time guarantees that the 
lower modes are not affected too strongly. Finally, if a, P, and y have the 

following expressions,- -< a < 0, /? =-, and y = — a, the HHT-a 

method is second-order accurate and unconditionally stable. With a = 0 the 
HHT-a method reduces to the constant acceleration method. The HHT-a 
method is useful in structural dynamics simulations incorporating many 
degrees of freedom (DOF), and in which it is desirable to numerically 
attenuate the response at high frequencies. Decreasing a below zero 
decreases the response at frequencies above provided that /? and y are 

defined as above. The procedure of the HHT-a method is summarized in 
Table F.l. 



ERDC/ITL TR-16-1 


Table F.l. HHT-a Method. 


1. Initial calculations 

1.1 Select At and a, (—^ < a < 0). 

1.2 Calculate B = (1 ~ a> and v = -- a. 

r 4 1 2 

1.3 Solve for it 0 mit 0 = (1 4- a)( p 0 - cit 0 - ku 0 ). 

1.4 Calculate c = (1 + a)c ••• k = (1 + a)k 

1.5 k = k 4- — c + 

/?At /?(At) 2 

1.6 a = — m + -c; and b = —m 4- At (— - 1) c. 

(3 At (3 ’ 2(3 \2 (3 ) 

2. Calculations for each time step, / 

2.1 c = ac(iii - Wj_i) k = ak{u t - u t _ x ) .*• / = (1 4- a)f i+1 -(14- 2a)/; 4- a/^. 

2.2 Apj = / + c + k ■■■ A pi = A pi + aiii + feiij. 

2.3 Solve for Ait; from kAu t = Ap;. 

2.4 Am i = -^—AUi -\u x 4- At (l - )it;. 

1 (3At 1 (3 1 V 2(3 J 1 

2.5 Ait/ = —A it/ —— ii, - —it,. 

1 (3 {At) 7 - 1 (3 At 1 2(3 1 

2.6 u i+1 = Uj + Au f , u i+1 = «i + Aiij, and u i+1 = it; + Ati,. 

3. Repetition for the next time step. Replace / by i+1 and implement steps 2.1 to 2.6 for the 
next time step. 




ERDC/ITL TR-16-1 


235 


Appendix G: Wilson-6 Method 

In Impact_Deck, the solution of the MDOF equations of motion is 
obtained by applying the Wilson-0 method. This method, developed by 
E.L. Wilson (2002), is a modification of the conditionally stable linear 
acceleration method that makes it unconditionally stable. This 
modification is based on the assumption that the acceleration varies 
linearly over an extended time step S t = 9 At, as shown in Figure G.i. The 
accuracy and stability properties of the method depend on the value of the 
parameter 0 , which is always greater than 1. 


Figure G.I Linear variation of acceleration over normal and 
extended time steps. 



The numerical procedure can be derived merely by rewriting the basic 
relationship of the linear acceleration method. The corresponding matrix 
equations that apply to MDOF systems are 

Au ; =(At)u ! . + ^Aii 1 . Au i =(At)u i + - ^ il^- J Ail { (G.I) 

Replacing At by St and the incremental responses by u t , Silt, and Sii as 
shown in Figure G.i gives the corresponding equations for the extended 
time step: 




















ERDC/ITL TR-16-1 


236 


c. /«. v . , 8f_. « /« \ (St ) 2 .. (St ) 2 

8u i = (St)u i +—8u i .-. 8u ; = (8f)u f H - u ; H - 8u ; (G.2) 


The right equation in Equation (G.2) can be solved for 


8a < = m Su ‘-&- 3ii < 


(G.3) 


Substituting Equation (G.3) into the left equation of Equation (G.2) gives 

Su i = ^ 8 u i- 3 u i~Y U - (G ' 4) 

Next, Equations (G.3) and (G.4) are substituted into the incremental (over 
the extended time step) equation of motion: 

mSii i + cSu i + k t 8u { = 8p t (G.5) 


where, based on the assumption that the exciting force vector also varies 
linearly over the extended time step. 


8p i .=efAp,.J 


This substitution leads to 


ki 8ui = 8 p { 


where: 


ki 



6 

(OA tf 


m 



0 


A Pi 


+ 


0 A t 


m + 3 c 


u ,+ 


0 0 Af 
3m H-c 


u- 


(G.6) 

(G.7) 

(G.8) 

(G.9) 


Equation (G.7) is solved for Sui, and Su t is computed from Equation (G.3). 
The incremental acceleration over the normal time step is then given by 



ERDC/ITL TR-16-1 


237 


Aii ; . =^8ii, (G.10) 

And the incremental velocity and displacement are determined from 
Equation (G.i). The procedure is summarized in Table G.i. 


Table G.I Wilson’s Method: Nonlinear Systems 


1. Initial calculations 

1.1 Solve mif'o = p 0 — cu 0 — (/ s ) 0 => u 0 . 

1.2 Select At and 9. 

1.3 a = -r-m + 3c; and b = 3m + —c. 

9At 2 

2. Calculations for each time step, / 

2.1 Spi = O(Api) + aiii + bill. 

2.2 Determine the tangent stiffness matrix k t . 

2.3 c + —m. 

2.4 Solve for 8u t from k t and 8p t 

2 -5 Su t = (eAf)2 Su t flAf u, 3 u t -, and Am, = g Su t . 

2.6 Au t = (A i)u t + y A ii t ; and Au t = (A t)u t + ^-ii t + ^-A ii t . 

2.7 u i+1 = Ui + Au if u i+1 = iii + Aii if and ii i+1 = u t 4- A ii t . 

3. Repetition for the next time step. Replace / by i+1 and implement steps 2.1 to 2.7 for the 
next time step. 


As mentioned earlier, the value of 0 governs the stability characteristics of 
Wilson’s method. If 0 = l, then this method reverts to the linear 
acceleration method, which is stable if At < 0.551 T N , where T N is the 
shortest natural period of the system. If 0 > 1.37, the Wilson -0 method is 
unconditionally stable, making it suitable for direct solution of the 
equations of motion and 0 = 1.42 gives optimal accuracy. 




ERDC/ITL TR-16-1 


238 


Appendix H: Member End Release Details for 
Load Applied at the End Release Node 

Member end release occurs in the Impact_Deck software when there is no 
moment transfer from one node to the next, because of an internal pin 
from one beam or deck to the next. Member end release details are 
discussed in Appendices A, B, and C for the generic case when the load is 
not applied at the end release node. 

When the load is applied at the end release node, an additional term is 
required for the end release equations and the load is altered to 
accomplish the lack of moment transfer. This Appendix extends the 
equations to include these terms. 

For a beam element, the six equilibrium equations in the local reference 
system can be written as 



1 , 2,...,6 


(H.l) 


If one end of the member has a hinge, or other type of release that causes 
the corresponding force to be equal to zero, Equation (H-i) requires 
modification. If we know that a specific value of f n is zero because of a 
release, the corresponding displacement u n can be written as: 


n—1 h- 12 Jc 

u = V-^ U.+ T — 

n ^ k J ^ h 

j= 1 #v nn j=n+ 1 r Sin 


U J+ r n 


(H.2) 


Therefore, by substitution of equation (H.2) into the other five equilibrium 
equations, the unknown u n can be eliminated and the corresponding row 
and column set to zero. Or: 


f ij = k ij u ij + r V 


(H.3) 


The terms ^ = r n = 0 and the new stiffness and load terms are equal to: 



ERDC/ITL TR-16-1 


239 


rt 


r-r — 

1 n k,„ 


(H.5) 


This procedure can be repeatedly applied to the element equilibrium 
equations for all releases. The repeated application of the simple 
numerical equation is sometimes called the static condensation or 
partial Gauss elimination. 



ERDC/ITL TR-16-1 


240 


Appendix I: Rayleigh Damping 

In this Appendix, the implementation of damping in the numerical 
procedures used in the computer software Impact_Deck will be discussed. 
Rayleigh damping is the method used. 

1.1 Method 

Consider first mass-proportional damping and stiffness-proportional 
damping: 


c = a Q m and c=a 1 k (i.i) 

Where the constants a 0 and % have units of sec -1 and sec, respectively. 
Physically, they represent the damping models shown in Figure I.i for a 
multistory building. Stiffness-proportional damping appeals to intuition 
because it can be interpreted to model the energy dissipation arising from 
story deformations. In contrast, mass-proportional damping is difficult to 
justify physically because the mass of the air damping the structure, 
compared to the structural mass, is negligibly smaller. Later we shall see 
that, by themselves, neither of the two damping models are appropriate 
for practical application. 

Figure 1.1 (a) Mass-proportional damping; (b) stiffness-proportional damping. 




(a) (b) 























ERDC/ITL TR-16-1 


241 


We now relate the modal damping ratios for a system with mass- 
proportional damping to the coefficient a 0 . The generalized damping for 
the n th mode, Equation (I.i) is 


C„ = a 0 M n 


( 1 . 2 ) 


and the modal damping ratio is 


(3 =^— 

I n o 

2 co„ 


( 1 - 3 ) 


The damping ratio is inversely proportional to the natural frequency. The 
coefficient a 0 can be selected to obtain a specified value of damping ratio in 
any one mode, say /?; for the i th mode. Equation (I.3) then gives 


a 0 = 2 Pi«i 


( 1 . 4 ) 


When a 0 has been computed, the damping matrix c can be determined 
from Equation (I.i). The damping ratio in any other mode, say the n th 
mode, is given by Equation ( 1 .3). Similarly, the modal damping ratios for a 
system with stiffness-proportional damping can be related to the 
coefficient a x . In this case 


c„ = a± co 2 n M n and p n 


( 1 . 5 ) 


The damping ratio increases linearly with the natural frequency. The 
coefficient a ± can be selected to obtain a specified value of the damping 
ratio in any one mode, say /?* for th ej th mode. Equation (I.5) then gives 


a. 


2 ?, 




( 1 . 6 ) 


When ay has been computed, the damping matrix c can be determined 
from Equation (I.i), and the damping ratio in any other mode is given by 
Equation (I.5). Neither of the damping matrices defined by the previous 
matrices is appropriate for practical analysis of MDOF systems. The 
variations of modal damping ratios with the natural frequencies they 



ERDC/ITL TR-16-1 


242 


represent are not consistent with experimental data indicating roughly the 
same damping ratios for several vibration modes of a structure. 

As a first step toward constructing a classical damping matrix somewhat 
consistent with experimental data, we consider Rayleigh damping 

c=a 0 m+a 1 k (1.7) 


The damping ratio for the n th mode of such a system is 


P 


a n 1 a 
——+—cd„ 
2 cd„ 2 " 


( 1 . 8 ) 


The coefficients a 0 and a 1 can be determined from specified damping 
ratios and /? ; for the i th and j th modes, respectively. Expressing 

Equation ( 1 . 8 ) for these two modes in matrix from leads to 


1 

2 



CD, 


CD. 


a A 


a. 



( 1 . 9 ) 


These two algebraic equation can be solved to determine the coefficients 
a 0 and a x . If both modes are assumed to have the same damping ratio /?, 
which is reasonable based on experimental data, then 


„ 2cd,cd, „ 2 

a 0 = p- l -^~ a 1 = P- 

CD; + CD, CD,. + CDj 


( 1 . 10 ) 


The damping matrix is then computed from Equation ( 1 .7) and the 
damping ratio for any other mode, given by Equation ( 1 . 8 ), varies with 
natural frequency. 

When applying this procedure to a practical problem, the modes i andj 
along with specified damping ratios should be chosen to ensure reasonable 
values for the damping ratios in all the modes contributing significantly to 
the response. 

In the Impact_Deck computer program this type of damping model is 
implemented to obtain the dynamic response. The user assigns a fixed 



ERDC/ITL TR-16-1 


243 


damping ratio and Impact_Deck estimates the natural frequencies and 
calculate the coefficients a 0 and a 1} and assemble the damping matrix of 
the system based on Equation (A. 31). 

1.2 Impact Deck frequency estimates 

For the model of an impact deck or beam directly mounted on a cluster of 
piles specified in Appendix A, Impact_Deck estimates the first two natural 
frequencies as follows: 


co 1 



( 1 . 11 ) 


where: 

k = spring transverse stiffness per unit length of the beam = (total 
number of springs * linear stiffness of one spring) / total length of 
the beam 

fh =the mass per unit length of the beam = mass density of the beam * 
cross-sectional area of the beam 

This estimate is based on SAP2000 models for a similar structure, and 
guidance from Chopra (2001). 

1.3 Guard Wall frequency estimates 

For the model of a guard wall consisting of simply supported beams 
specified in Appendix B, Impact_Deck estimates the first two natural 
frequencies as follows, 


«i = 


2 . 5 * k 


m 


.\cd 2 = 


3.0 *k 


m 


( 1 . 12 ) 


where: 

k =the transverse elastic stiffness of the center pile group 
m =the total mass of the two beams = mass density of the beam * 
cross-sectional area of the beam * total length of the structure 


This estimate is based on SAP2000 models for a similar structure. 



ERDC/ITL TR-16-1 


244 


1.4 Flexible Wall frequency estimates 

For the flexible wall problem specified in Appendix C, Impact_Deck 
estimate the first two natural frequencies as follows: 


«i 



1 2.0 *m 


( 1 . 13 ) 


where: 

k =the transverse elastic stiffness of the center pile group 
m =total mass of one beam = mass density of the beam * cross- 
sectional area of the beam * length of one beam 
m* =total mass of the two beams = mass density of the beam * cross- 
sectional area of the beam * total length of the system including 
the length of the shear key. 

This estimate is based on SAP2000 models for a similar structure. 



ERDC/ITL TR-16-1 


245 


Appendix J: Key Impact.Deck Program 
Variables 

NSTU mode of operation: 

1. Flexible Wall 

2. Guard Wall 

3. Impact Deck 

DS<X,Y,R>() Input displacements (fixed) 

FS <X,Y,R> 0 Input force/moment (fixed) 

These variables describe a user-defined, piece-wise linear, force 
displacement curve for piles. 

SPRING_K<X,Y,R>() Computed tangent stiffness of a pile group 

SPRING_B<X,Y,R>() Computed force axis intercept a pile group 

These variables return the tangent stiffness and force intercept 
(deflection=o) at each part of the user-defined, piece-wise linear, force 
displacement curve described by DS<X,Y,R>() and FS<X,Y,R>(). By 
working from the force intercept, the appropriate force for an absolute 
displacement can be computed 

SK_SPRING() Current stiffness tangent at pile locations 

SB_SPRING() Current stiffness force axis intercept at pile 

locations 

Returned from the stif¥hess() function, these values return the current 
tangent stiffness and intercept that the pile is under. This value takes into 
considerations whether the pile is under an unload/reload cycle where the 
resisting force is less than the maximum force achieved to that point. 

NELASQ Flag for elastic loading condition: 


1. Load 



ERDC/ITL TR-16-1 


246 


2. Unload/Reload 


uelas<x,y,R> 


X_LOAD() 


Greatest deflection to this point - for 
unload/reload cycle 

The x-position of the applied load over time 
(along the wall) 


This is pre-computed as a linear increment from the start point to the end 
point of the simulation, given in block 4 of the input file 


U(,) 

UDC) 

UDD(,) 

SK(,) 

SM(,) 

SKM(,) 

SMMQ 

SKMMC) 

SKMMS(,) 

SMMM(,) 


The absolute displacement of each pile over 
time 

The absolute velocity of each pile over time 

The absolute acceleration of each pile over time 

Element stiffness matrix 

Element mass matrix 

Global stiffness matrix 

Global mass matrix 

The global stiffness matrix (sans restraints on 
DOF) 

Intermediate matrix multi value 

The global mass matrix (sans restraints on 
DOF) 


SCMMC) 

ALFAA 

BETAA 


The global damping matrix 
Rayleigh damping term for mass - ao 
Rayleigh damping term for stiffness - ai 



ERDC/ITL TR-16-1 


247 


These variables are determined from the natural frequencies of the 
structure and □ (EFIl in Block 5 of the Input File) 


P_LOAD(,) Force acting at nodes of the element at each 

time step (spread to adjacent nodes to the 
impact) 

FOR_SPRING(,) Element spring forces occurring at each node 

at each time step 


ELENGTHQ 


Length of each wall segment (in the X direction 
only) 


NRES 


Total number of array indexes (~= number of 
nodes*number of DOF) 


NODF<i, 2,3>() Index for DOF for node number 


NOND 


Number of nodes 


NOTE: may not be the number of pile groups 

NPS() Indices for DOF information for pile groups 

(Flexible and Guard Walls) 



REPORT DOCUMENTATION PAGE 

Form Approved 

OMB No. 0704-0188 

Public reporting burden for this collection of information is estimated to average 1 hour per response, including the time for reviewing instructions, searching existing data sources, gathering and maintaining 
the data needed, and completing and reviewing this collection of information. Send comments regarding this burden estimate or any other aspect of this collection of information, including suggestions for 
reducing this burden to Department of Defense, Washington Headquarters Services, Directorate for Information Operations and Reports (0704-0188), 1215 Jefferson Davis Highway, Suite 1204, Arlington, VA 
22202-4302. Respondents should be aware that notwithstanding any other provision of law, no person shall be subject to any penalty for failing to comply with a collection of information if it does not display a 
currently valid OMB control number. PLEASE DO NOT RETURN YOUR FORM TO THE ABOVE ADDRESS. 

1. REPORT DATE (DD-MM-YYYY) 2. REPORT TYPE 

July 2016 

3. DATES COVERED (From - To) 

4. TITLE AND SUBTITLE 

Simplified Dynamic Structural Time-History Response Analysis of Flexible Approach 

Walls Founded on Clustered Pile Groups Using Impact_Deck 

5a. CONTRACT NUMBER 

5b. GRANT NUMBER 

5c. PROGRAM ELEMENT NUMBER 

6. AUTHOR(S) 

Barry C. White, Jose Ramon Arroyo, and Robert M. Ebeling 

5d. PROJECT NUMBER 

5e. TASK NUMBER 

5f. WORK UNIT NUMBER 

448769 

7. PERFORMING ORGANIZATION NAME(S) AND ADDRESS(ES) 

Information Technology Laboratory 

U.S. Army Engineer Research and Development Center 

3909 Halls Ferry Road, Vicksburg, MS 39180-6199; 

Department of General Engineering 

University of Puerto Rico, Mayaguez, PR 00681 

8. PERFORMING ORGANIZATION REPORT 
NUMBER 

ERDC/ITL TR-16-1 

9. SPONSORING / MONITORING AGENCY NAME(S) AND ADDRESS(ES) 

U.S. Army Corps of Engineers 

441 G. Street NW 

Washington, DC 20314-1000 

10. SPONSOR/MONITOR’S ACRONYM(S) 

11. SPONSOR/MONITOR’S REPORT 
NUMBER(S) 


12. DISTRIBUTION / AVAILABILITY STATEMENT 

Approved for public release; distribution is unlimited. 


13. SUPPLEMENTARY NOTES 


14. ABSTRACT 

Flexible approach walls are being considered for retrofits, replacements, or upgrades to Corps lock structures that have exceeded their 
economic lifetime. This report discusses a new engineering software tool to be used in the design or evaluation of flexible approach 
walls founded on clustered pile groups and subjected to barge train impact events. 

This software tool, Impact_Deck, is used to perform a dynamic, time-domain analysis of three different types of pile-founded flexible 
approach walls: an impact deck, an alternative flexible approach wall, and a guard wall. Dynamic loading is performed using impact- 
force time histories (Ebeling et al. 2010). This report covers the numerical methods used to create this tool, a discussion of the graphical 
user interface for the tool, and an analysis of results for the three wall systems. 

The results of analyzing the three wall systems reveals that dynamic evaluations should be performed for these structures because of 
inertial effects occurring in the wall superstructure and substructure. These inertial effects can cause overall and individual response 
forces that are greater than the peak force from the impact-force time history. 


This report also discusses the advantages of load sharing between multiple pile groups in an approach wall substructure. In the case of 
Lock and Dam 3, the peak reaction force for any individual pile group was 11% of the peak impact load. 


15. SUBJECT TERMS 

See reverse 

16. SECURITY CLASSIFICATION OF: 


17. LIMITATION 

18. NUMBER 

19a. NAME OF RESPONSIBLE 




OF ABSTRACT 

OF PAGES 

PERSON 

a. REPORT 

b. ABSTRACT 

c. THIS PAGE 



19b. TELEPHONE NUMBER (include 

Unclassified 

Unclassified 

Unclassified 


264 

area code) 


Standard Form 298 (Rev. 8-98) 

Prescribed by ANSI Std. 239.18 















15. SUBJECT TERMS (concluded) 

Barge impact 
Barge train impact 
Flexible lock approach wall 
ImpactDeck 
ImpactForce 
Flexible approach wall 
Guide wall 
Guard wall 
Pile groups 
Clustered pile groups 
Glancing blow 
Impact Deck 
Time-history analysis 
Force time-history 
Dynamic analysis 
Dynamic structural analysis 
Simply supported beam