Вы находитесь на странице: 1из 114

Report No.

TR-97-2

TECHNICAL REPORT

FOUNDATION GUIDELINES

By
A. J. Smalley
P. J. Pantermuehl

Mechanical and Fluids Engineering Division


Southwest Research Institute

January 1997

© 1998 Gas Machinery Research Council


This document contains information resulting from a cooperative research
effort. The contents hereof are only intended to be guidelines for the subject
matter to which the document pertains. Neither Southern Gas Association
nor the Gas Machinery Research Council make any warranty or
representation, express or implied, with respect to the accuracy,
completeness or usefulness of the information contained in this document,
including, without limitation, implied warranties of merchantability and
fitness for a particular purpose, or that the use of any method, suggestion,
technology, information or guidelines disclosed herein may not infringe on
rights owned or claimed by others. In no event will Southern Gas
Association or the Gas Machinery Research Council be liable for any
damages, including, without limitation, liability arising out of contract,
negligence, strict liability, environmental or tort, warranty or copyright
infringement, or any incidental or consequential damage arising out of the
use of this Report. The user assumes any liability with respect to any
methods, suggestions, technology, guidelines or other information
contained herein and releases Southern Gas Association and the Gas
Machinery Research Council from any and all damage, loss or injury having
to do with use of any such methods, suggestions, technology, guidelines or
other such information.

This document may contain references to product(s) which may assist in


achieving one or more guidelines as may be set forth herein. Such
references are not intended to constitute endorsement or criticism of any
such product(s) by the Gas Machinery Research Council or Southwest
Research Institute. Any attempted use of this Report, or its contents, by
anyone, as an endorsement or criticism of any such product(s) is expressly
prohibited. Neither this Report or its contents may be used for any
advertising purposes whatsoever.

GMRC PURPOSE

The Gas Machinery Research Council provides member companies and industry with the
benefits of an applied research and technology program directed toward improving
reliability and cost effectiveness of the design, construction, and operation of mechanical
and fluid systems.

For additional copies of this report, please contact:

Marsha Short
Director, Member Services
Gas Machinery Research Council
3030 LBJ Freeway, Suite 1300, L.B. 60
Dallas, TX 75234
Telephone (972) 620-4024
FAX (972) 620-8518
EXECUTIVE SUMMARY

These guidelines provide guidance to engineers responsible for installing or repairing the
foundations of reciprocating compressors. If followed, the guidance should help assure an
installation with a long life of trouble-free service. They address, in particular, mechanical
integrity of the compressor/foundation installation, with emphasis on design of mounts, anchor
bolts, and foundation block to carry the required dynamic loads. They defer to existing texts on
vibration engineering for foundation blocks on their soil.

The guidelines emphasize the importance of considering forces from individual cylinders and
throws when engineering mounts, anchor bolts, and foundation block internal stresses. The
summed force and moments typically provided by the OEM result from summing individual
cylinder forces and moment contributions; they usually mask the magnitude of individual forces
which individual mounts must carry. These summed forces are normally provided for use in
vibration analysis of the compressor and block on soil; however, they are generally inappropriate
for determining how to hold the compressor to its foundation.

The introduction to the guidelines defines scope and responsibilities. Subsequent sections address
forces generated within the compressor; the transmission of these forces through the frame to
mounts and block; compressor mounting; anchor bolts; and the foundation block itself, including
internal stresses, and strength considerations.

The guidelines include brief sections on installation; operations and maintenance; inspection and
condition monitoring.

Installation is critical to the accomplishment of the intent of the guidelines, but depends heavily on
an experienced installer; the guidelines do not purport to replace that experience, but the section on
Installation identifies a few relevant topics within the guidelines.

The section on O&M touches on maintenance practices which can help ensure the long-term
integrity of the installation.

The section on Condition Monitoring and Inspection lists practices which will help identify
problems which need to be addressed during the life of a foundation and the compressor’s
mounting system.

ii
The guidelines include an additional section (Summary of Guidelines), which in 46 individual one-
line items summarize the key points of the guidelines for easy reference.

The appendices to the guidelines include the calculation of unbalanced shaking forces and
moments; the summary of a finite element analysis of a reciprocating compressor on its foundation;
results of some PCRC pull-out tests on low and high strength anchor bolts; and a glossary.

iii
TABLE OF CONTENTS

Page

1.0 INTRODUCTION 1-1


1.1 General 1-1
1.2 Scope and Limitations 1-1
1.2.1 Machinery Considered 1-1
1.2.2 Issues Addressed 1-2
1.3 Foundation Requirements 1-2
1.3.1 Function 1-2
1.3.2 Life 1-2
1.3.3 Life Cycle Considerations 1-3
1.4 Responsibilities 1-3
1.4.1 End User 1-3
1.4.1.1 Project Engineer 1-3
1.4.1.2 Operations and Maintenance Staff 1-3
1.4.2 Equipment Supplier 1-4
1.4.3 Designer 1-4
1.4.4 Installer 1-4
1.4.5 Communication 1-4
1.5 Summary of the Guidelines’ Role 1-5
1.6 Caution 1-5

2.0 FORCES FROM THE COMPRESSOR AND DRIVER 2-1


2.1 Compressor Function and Configuration 2-1
2.1.1 Cylinder and Piston Action 2-1
2.1.2 Multiple Stages 2-1
2.2 Compressor Gas Forces 2-2
2.2.1 General 2-2
2.2.2 Time Varying Gas Forces 2-2
2.2.3 Maximum Gas Forces 2-3
2.2.4 Accounting for Non-Ideal Influences 2-3

iv
TABLE OF CONTENTS
(continued)
Page

2.0 FORCES FROM THE COMPRESSOR AND DRIVER (cont’d)


2.3 Compressor Inertia Forces 2-4
2.3.1 General 2-4
2.3.2 Time Varying Inertia Forces 2-4
2.3.3 Maximum Inertia Forces 2-6
2.4 Summed Unbalanced Forces and Moments 2-6
2.4.1 General 2-6
2.4.2 Eight Force and Moment Component Definition 2-7
2.4.3 Calculating Summed Forces and Moments 2-7
2.5 Separable Drive Torques and Forces 2-7
2.6 Managing Inertia Forces and Moments 2-8
2.6.1 Global Considerations Based on Summed Forces and Moments 2-8
2.6.2 Local Considerations 2-9
2.7 Managing Local Gas Forces 2-9

3.0 TRANSMISSION OF FORCES 3-1


3.1 Overview and Structural Considerations 3-1
3.2 Gas Force Transmission 3-1
3.3 Inertia Force Transmission 3-2
3.3.1 How Flexibility Impacts Local Forces 3-2
3.3.2 Measured Flexibility 3-2
3.4 Calculation Methods 3-2
3.4.1 Conservative Calculation Method 3-2
3.4.2 Non-Conservative Calculation Method 3-3
3.4.3 Finite Element Method 3-3

v
TABLE OF CONTENTS
(continued)

Page

3.0 TRANSMISSION OF FORCES (cont’d)


3.4.4 Guidance from Previous Finite Element Analysis 3-3
3.4.4.1 Basis for Guidance 3-3
3.4.4.2 Calculating Horizontal Forces to be Restrained (General) 3-3
3.4.4.3 Forces Transmitted by Crosshead Guide 3-4
3.4.4.4 Required Tie-Down Capacity at CHG Bolts 3-4
3.4.4.5 Required Tie-Down Capacity at Frame Bolts 3-4
3.4.4.6 Safe Force Reduction Factors 3-4
3.4.4.7 Summary of Force Formulae 3-5

4.0 COMPRESSOR MOUNTING 4-1


4.1 Overview and Requirements 4-1
4.2 The Role of Friction 4-1
4.2.1 Relating Tangential and Normal Forces 4-1
4.2.2 Anchor Bolt Requirement 4-1
4.2.3 Friction Data 4-2
4.3 Chock Configuration 4-2
4.3.1 Benefits of Chock Mounting 4-2
4.3.2 Steel’s Advantages 4-3
4.3.3 Steel’s Disadvantages 4-3
4.3.4 Epoxy’s Advantages 4-3
4.3.5 Epoxy’s Disadvantages 4-3
4.3.6 Chock Design 4-3
4.3.7 Normal Load Limits 4-3
4.3.8 Effects of Creep and Anchor Bolt Length 4-4
4.4 Anchor Bolt Preload 4-5

vi
TABLE OF CONTENTS
(continued)

Page

4.0 COMPRESSOR MOUNTING (cont’d)


4.5 Thermal Distortion and Growth 4-6
4.5.1 Block Distortion 4-6
4.5.2 Frame Distortion 4-6
4.5.3 Cyclic Frame Distortion 4-6
4.5.4 Alignment Implications of Thermal Distortion 4-7
4.5.5 Weight Load Implications of Thermal Distortion 4-7
4.5.6 Horizontal Load Implications of Thermal Growth 4-7

5.0 ANCHOR BOLTS 5-1


5.1 Material 5-1
5.1.1 High Strength Material for New Installations 5-1
5.1.2 Lower Strength Bolts in Old Installations 5-1
5.2 Anchor Bolt Diameter 5-1
5.3 Threads 5-1
5.4 Length 5-2
5.5 Configuration 5-3
5.5.1 Avoid J and L Bolts 5-3
5.5.2 Symmetrical Terminations 5-3
5.5.3 Square or Rectangular Terminations 5-3
5.5.4 Large Termination Plates 5-3
5.5.5 Nut Terminations 5-4
5.5.6 Through to the Mat Configuration 5-4
5.5.7 Canister Bolt 5-4
5.5.8 Supplementary Strength of Old Anchor Bolts 5-4
5.6 Preload 5-5
5.7 Sleeving 5-5
5.7.1 Relief Near Top of Bolt 5-5
5.7.2 Bolt Stretch Sleeving 5-5

vii
TABLE OF CONTENTS
(continued)

Page

5.0 ANCHOR BOLTS (cont’d)


5.8 Nuts and Washers 5-6
5.8.1 Spherical Self-Aligning Washers 5-6
5.8.2 Hex Nuts 5-6
5.8.3 Super Nuts 5-6
5.9 Anchorage 5-7
5.9.1 Strength Experiences 5-7
5.9.2 Design of Anchorage 5-7

6.0 FOUNDATION BLOCK 6-1


6.1 Overview 6-1
6.2 Concrete 6-1
6.2.1 Positive Characteristics for Compressor Foundations 6-1
6.2.2 Tensile Strength of Concrete 6-1
6.2.3 Evolving Characteristics 6-2
6.3 Rebar 6-2
6.4 Foundation Geometry 6-3
6.4.1 Shape 6-3
6.4.2 Stress Concentration 6-3
6.4.3 Reducing Stress Concentration in the Pan Recess 6-3
6.4.4 Controlling Nominal Tensile Stress 6-3
6.4.5 Interfaces 6-3
6.4.6 Location of Anchor Bolts 6-4
6.5 Loads and Concrete Stresses 6-4
6.5.1 General 6-4
6.5.2 Termination Loads 6-4
6.5.3 Horizontal Loads and Stresses 6-4
6.6 Calculation and Assessment of Stresses 6-5
6.6.1 Finite Element Methods 6-5
6.6.2 Bolt Termination Stresses 6-5
6.6.3 Estimating Pan Recess Stresses 6-6

viii
TABLE OF CONTENTS
(continued)

Page

6.0 FOUNDATION BLOCK (cont’d)


6.7 Post-Tensioning 6-8

7.0 INSTALLATION 7-1

8.0 OPERATION AND MAINTENANCE 8-1

9.0 INSPECTION AND CONDITION MONITORING 9-1

1 0 . 0 SUMMARY OF GUIDELINES 10-1

1 1 . 0 REFERENCES 11-1

1 2 . 0 ACKNOWLEDGMENTS 12-1

APPENDIX A - CALCULATION OF UNBALANCED SHAKING


FORCES AND MOMENTS A-1
The Development of the Computational Method A-2
Analysis for Counterweights A-9

APPENDIX B - SUMMARY OF FINITE ELEMENT ANALYSIS OF A


RECIPROCATING COMPRESSOR AND ITS FOUNDATION B-1
Modeling of Compressor Foundation Interaction B-2
Influence of Frame Flexibility on Transmitted Forces B-3
Response of the Concrete Block to Applied Loads B-5

APPENDIX C - FIELD DATA SUMMARY REPORT ON PULL-OUT


TESTS C-1

APPENDIX D - GLOSSARY D-1

ix
LIST OF FIGURES

Figure Page

2-1a Plan View of Driver Connected to Driven Machine 2-11

2-1b Free Body Diagram in Plan View, Showing Torques from 2-11
Attached Machine and Reaction Torques from Foundation

2-1c Free Body Diagram in End Elevation Looking from Driver End 2-12
Towards Driven Machine Showing Torques from Attached
Machine, Reaction Torques from Foundation, and Torques
Applied to Foundation

2-1d Free Body Diagram from a Frame in Which Drive Torque is 2-12
Reacted by Different Net Forces on Either Side of the Frame

3-1 Deflected Shapes of a Compressor Frame for Each 45 Degrees of 3-6


Crankshaft Rotation

3-2 Influence of Assumptions on Forces to Block, CHG Bolts, and 3-6


Frame Bolts

4-1 Types of Compressor Frame Support Systems 4-9


(Illustrations Courtesy of Robt. L. Rowan & Assoc., Inc.)

4-2 Adjustable Chocks 4-10

4-3 Full Bed Grout 4-11

5-1 Evolution of Anchor Bolt Designs 5-8


(Illustration Courtesy of Robt. L. Rowan & Assoc., Inc.)

5-2a Multiple Termination Point Anchor Bolt Supplement 5-9

5-2b Multiple Extension Method of Increasing Anchorage Strength for 5-10


Old Anchor Bolts
(Illustration Courtesy of Robt. L. Rowan & Assoc., Inc.)

5-3 Multi Jackbolt Mechanical Tensioner Installation on Crosshead 5-11


Guide Tie-Down

6-1 Foundation Block Isometric View - Compressor Area Showing 6-10


Grout Pocket, Sole Plates and Observed Areas of Cracking

6-2 Predicted Influence of Horizontal and Vertical Bolts on Tendency 6-10


of Concrete to Crack in Tension Near Pan

6-3 Configuration for Estimating Stress in the Pan Area 6-11

x
LIST OF FIGURES
(continued)

Figure Page

6-4a Example Block with Lateral and Longitudinal Post-Tensioning 6-12


Rods
(Illustrations Courtesy of PanEnergy Corp.)

6-4b Termination Details for Post-Tensioning Rods 6-13


(Illustrations Courtesy of PanEnergy Corp.)

A-1 Cylinder Geometry A-11

B-1 Model of Block and Frame B-7

B-2 Maximum Gas Force Rod Load for 3 Stage Compressor B-7

B-3 Variation of Gas Forces for Cylinder #5 B-8

B-4 Comparison of Gas Load to Crosshead Guide Loads Based on B-8


Flexible Frame (FE) Analysis & Rigid Frame Assumption

B-5 Transmissibility Effects: Comparison of Inertia Load on B-9


Crankshaft to Frame Bolt Loads Predicted for Flexible
Block/Frame & Rigid Frame

B-6 Anchor Bolt Loads for Rigid, Flexible, & Soft Frames B-9

B-7 Comparison of Rigid Frame, Flexible Frame, and Realistic Forces B-10
for Reciprocating Compressor Foundations

B-8 Model of Foundation Block for Large Process Compressor B-10

B-9 Predicted Extent of Cracking at Anchor Bolt Termination B-11

B-10 Tensile Stress Contours in Pan Recess B-12

C-1 Test Set-up - Anchor Bolt Pullout C-3

C-2 Block 1 (Low Strength, Test Bolt 1) C-4

C-3 Block 1 (Low Strength, Test Bolt 3) C-5

C-4 Block 1 (Low Strength, Test Bolt 5) C-6

C-5 Block 1 (Low Strength, Test Bolt 6) C-7

C-6 Block 4 (Low Strength, Bolt 1) C-8

C-7 Block 2 (High Strength Bolt) C-9

xi
LIST OF TABLES

Table Page

6-1 Values for Stress Concentration Factor, Ksc, Inferred from 6-9
Reference 17

B-1 Summary Influence of Modeling Assumption Compared to B-4


Conservative Assumptions for Prediction Forces

xii
1.0 INTRODUCTION

1.1 General

These guidelines address the mounting of reciprocating compressors on concrete block foundations
with the goal of assuring a sound installation and a long life of productive service.

Sources include discussion with experts and specific analyses supplemented by observation and
judgment. The guidelines seek to maximize use of detailed modeling recently undertaken by the
PCRC, to maximize use of past research by PCRC, PRCI, and others (Refs. 1 through 9), and to
fill gaps with consensus experience and judgment; where potential for disagreement exists, to
provide fundamental objectives, and alternative ways to accomplish these. Knowledge on this
subject remains incomplete, but PCRC’s ongoing support of relevant analysis and test defines a
process of continuous improvement which should supplement and consolidate current knowledge,
and experience.

As discussed below under the heading of Responsibilities, some important elements of achieving a
satisfactory installation include:

• Recognition that the compressor and foundation must form an integrated structure.
• Recognition of the installation as a system with interacting interfaces.
• Communication between parties with shared responsibilities.
• Understanding of how each responsibility contributes to the whole, and
• Awareness of how each responsibility overlaps that of other parties.

The PCRC has sought to assemble guidelines which operating companies can adapt to their
individual specifications.

1.2 Scope and Limitations

1 . 2 . 1 Machinery Considered

The guidelines address, most directly, large low speed compressors mounted permanently on
concrete block foundations. The compressor may contain an engine driver within the same frame;
alternatively a separate motor, reciprocating engine, or turbine may drive the reciprocating

1-1
compressor. For turbine drives, a gear box normally reduces turbine speed to the compressor's
low speed. Motors most commonly drive directly. The guidelines attempt to focus on features
special to low speed reciprocating compressors with horizontal compressor cylinders. The term
low speed does not have a precise definition, but normally covers the range 200 RPM to 600
RPM.

While information may also guide those installing high speed units or skid mounted units, the
guidelines have not sought generality for such installations.

1 . 2 . 2 Issues Addressed

The scope concentrates on mechanical integrity of the compressor structure, its mounting, and the
concrete foundation block. It does not attempt to address interaction of foundation and soil. While
recognizing the importance of controlling natural frequencies of the block upon the soil, of
avoiding resonance by providing adequate margin between these natural frequencies and first or
second order of running speed, and of limiting vibration amplitude of the compressor on its
foundation, the guidelines leave this subject and methods of accomplishing these essential goals to
other texts (Refs. 10 through 12).

1.3 Foundation Requirements

1 . 3 . 1 Function

The foundation for a reciprocating compressor installation must position and support the
compressor. It must keep the compressor aligned and help it support the loads generated by the
compressor and its driver. In supporting time varying loads, the combined stiffness of the
compressor’s structure and foundation must limit dynamic motion of the compressor and bending
loads on the crankshaft to acceptable levels. In summary, the compressor and its foundation must
act together as an integrated structure which assures sound, reliable, robust machine function.

1 . 3 . 2 Life

The foundation should meet these requirements over the anticipated life of the compressor
installation. It should, therefore, support all loads which it experiences with sustained strength and
integrity. The stresses which the loads generate within the block should fall below the levels at

1-2
which cracking may occur for the concrete in question. Concrete strength under dynamic loads
should consider fatigue and the reduction in strength it causes. If cracking occurs, it should remain
localized, with its growth limited.

1 . 3 . 3 Life Cycle Considerations

Some of these requirements demand attention in installation, operation, maintenance, condition


assessment, and repair for the compressor and its foundation, as well as in the up front design.

Many compressors have run for 30, 40, 50 years and even longer; mostly their original
foundations eventually deteriorate, and require repair. The repair work involved will range from
replacement of an 18-inch layer of original concrete block to complete replacement of the
foundation from the mat up. The repaired foundation and mounting system also needs to meet the
above requirements.

1.4 Responsibilities

1 . 4 . 1 End U s e r

The end user of the equipment carries the ultimate responsibility for assuring the foundation
meets the requirements, since the consequences of an unsatisfactory installation directly influence
the ability of the equipment to generate revenue for the end user.

1.4.1.1 Project Engineer

Within the end user organization the project engineer should ensure a satisfactory, robust
design and installation. This includes ensuring that suppliers, designers, and installers
communicate, provide needed information to other parties, and meet their responsibilities.

1.4.1.2 Operations and Maintenance Staff

While robustness implies some tolerance to operational variances, the operator and
maintenance staff will benefit from awareness of the most sensitive aspects of the installation,
which normally include the mounting interfaces between compressor and concrete, sensitivity to
the uncontrolled presence of oil on the structure, the need to keep anchor bolts tight, and a good
inspection program.

1-3
1 . 4 . 2 Equipment Supplier

For new equipment, the equipment supplier (compressor OEM) carries a responsibility to
provide some important pieces of data and information, without which these assurances cannot be
met. The required information may go beyond what equipment suppliers have traditionally
provided, particularly forces from individual throws and cylinders.

1 . 4 . 3 Designer

The designer of the foundation and compressor mounting system carries the
responsibility to obtain the required data, and to apply the data and information effectively so the
foundation meets its requirements.

1 . 4 . 4 Installer

The installer of the compressor and foundation must carry out the intent of the design and apply
experience to achieve in the installation process the integrity inherent in the design. The installer
should understand the intent and significance of all design features and specifications. The installer
must manage details such as the aggregate used in the concrete, control of the mix from batch plant
to pour, any special requirements of high and early strength concrete, the curing process of
concrete and grouts, surface coatings, expansion joints, templates for bolt positioning, leveling of
the soleplates, alignment of the compressor, appropriate tightening and loosening sequences for
frame tie-bolts, and for anchor bolts. While the guidelines attempt to consider as many important
issues as possible, some of these subjects depend on experience not readily duplicated by a
document such as this. The subject of grouts have received detailed attention by the ACI in their
Report ACI 351.1R-93 (Ref. 13), and the reader should take advantage of this and other relevant
references.

1 . 4 . 5 Communication

All responsible parties must communicate effectively, using common terms, with common
meanings, and must share an understanding of the needed information and how it influences the
process of achieving a sound installation. A compressor, its mounts, and its foundation act as a
system with physical interfaces. Those with responsibilities for it have organizational interfaces.
There exists strong potential for miscommunication, or for the assumption that another party

1-4
carries entirely what may be an overlapping responsibility. Design features of the foundation and
mount may influence some of the information expected from the manufacturer, and design features
of the compressor may limit options open to the foundation designer. The following guidelines
seek to assist the process, and to identify as clearly as possible how to accomplish the assurances
summarized below.

1.5 Summary of the Guidelines’ Role

In summary, the guidelines should help assure:

• Awareness of forces which the compressor applies to mounts and block.


• Satisfactory management of these forces.
• Reliable transmittal of dynamic forces at the mount interfaces without slippage.
• Integrity of the mount (chock/rail/soleplate) under loads from the anchor bolt and friction.
• Anchor bolts which provide needed interface forces without compromising block integrity.
• A block and mat configured to carry the horizontal and vertical loads without damage.
• Understanding of maintenance needs (including tight anchor bolts at all times).
• Understanding of inspection needs for identifying developing problems.
• Use of common terms (see Glossary).

1.6 Caution

The guidelines contain reference with acknowledgment to certain commercial


products which may help accomplish the guidelines. These references do not
represent an endorsement by PCRC, nor that PCRC has performed any
independent tests of their performance, nor that these are the only available
products; a more extensive search could find others that may meet the need.
Some products are patented. As good practice, any potential user of such a
product should seek from the supplier or by independent test, data confirming its
performance and assurance of the likely effectiveness of the product in the
planned application.

1-5
2.0 FORCES FROM THE COMPRESSOR AND DRIVER

2.1 Compressor Function and Configuration

2 . 1 . 1 Cylinder and Piston Action

A reciprocating compressor raises the pressure of a certain flow of gas by imposing reciprocating
motion on a piston within a cylinder. The piston normally compresses gas during both directions
of reciprocating motion. To accomplish this double-acting compression, each cylinder comprises
two “ends” within each of which piston motion causes the volume to vary. Gas flows to and from
each end through passive spring loaded valves (“check valves”); one valve allows gas to flow
into the cylinder from a low pressure source (the suction valve), and one allows gas to flow out of
the cylinder when its pressure exceeds the pressure within a discharge manifold (the discharge
valve).

A crankshaft, usually with a number of throws, moves each piston via a slider-crank mechanism.
More specifically, a crosshead driven by the connecting rod, slides horizontally on slipper bearings
(mounted in a crosshead guide) and imparts this horizontal motion to the piston. The cylinder end
away from the crankshaft is referred to as the head end or the outboard end. The cylinder end
nearest the crankshaft is referred to as the crank end or the inboard end. Most commonly, the
active (swept) area of the head end, on which the head end pressure acts is the full cylinder bore,
and the swept area of the crank end is reduced from the full cylinder bore by the area of the piston
rod (which applies the compression load to the piston). In some high pressure applications, the
head end may employ a tail rod, in which case the head end swept area is also reduced by the tail
rod area. In some configurations, a second (“tandem”) cylinder is located outboard of the first
cylinder. Here the inboard piston has an outer rod attached which drives the piston of the outboard
cylinder; and again, the head end with the rod reduces the swept area on which cylinder pressure
acts.

2 . 1 . 2 Multiple Stages

The compressor may have one or more stages of compression. For a single stage, all cylinders see
similar values for suction and discharge pressure. For two or more stages, the gas discharged
from one stage become suction gas for the next stage, typically after intercooling, and some loss in

2-1
pressure between stages. In most cases, the flow from one stage equals the flow to the next stage,
but exceptions occur with side streams to or from interstages, with liquid removal between stages,
or with the use of a common header for interstages of several parallel compressors.

A configuration occasionally encountered accomplishes two stages of compression in a single


cylinder, using the head end for one stage, and the crank end for another. In this case, the suction
and discharge pressures for the two ends differ.

2.2 Compressor Gas Forces

2 . 2 . 1 General

The following formulation develops forces acting within individual cylinder and on individual
rods. These individual cylinder forces have direct relevance to the calculation of forces transmitted
to crosshead tie-downs, mounts, and foundation.

2 . 2 . 2 Time Varying Gas Forces

In a double-acting cylinder, the instantaneous net compressive gas force contributed to the piston
rod equals the difference between head and crank end gas forces:

Frod = Phead*Ahead-Pcrank*Acrank (2-1)

where:

2 2
Ahead or Acrank = (π/4)*(B -k*Drod )
k = one or zero according to whether the end has a rod or not.
B = the cylinder bore.
Drod = the rod diameter.
P head, Pcrank = the instantaneous head and crank pressures.

The existence of a second (“tandem”) cylinder on the same rod would add a gas force contribution
based on the head and crank pressures and areas for that cylinder.

2-2
2 . 2 . 3 Maximum Gas Forces

The head and crank end pressures vary continuously and the differential force takes both positive
and negative (inward and outward) net values during each cycle of piston motion. Assuming the
rod load reverses for at least part of the cycle, for design purposes, we need the highest values of
compressive and tensile rod force:

(Frod) gc-max = Max (Phead * Ahead - Pcrank * Acrank) (2-2)


(Frod) gt-max = Max (Pcrank * Acrank - Phead * Ahead) (2-3)

A widely used approach to establishing these quantities sets the minimum and maximum values for
pressure in each end to be suction and discharge pressure for the stage in question, then multiplies
the result by a factor, F1:

(Frod)gc-max = (Pdhead*Ahead-Pscrank*Acrank) * F1 (2-4)


(Frod)gt-max = (Pdcrank*Acrank-Pshead*Ahead) * F1 (2-5)

where Pshead, Pdhead are the suction and discharge pressures which the head end sees, P scrank, Pdcrank
are the suction and discharge pressures which the crank end sees.

2 . 2 . 4 Accounting for Non-Ideal Influences

The factor F1 helps account for flow resistance in the valves and outside the cylinder which forces
maximum pressure above discharge pressure, and minimum pressure below suction pressure to
maintain flow into and out of the cylinder, so raising the maximum differential across the piston.
In addition, F 1 addresses pulsations in the piping and nozzles which tend to cause modulation in
the pressure outside the cylinder, and to further raise the maximum differential.

By setting F1 to 1.0, we base rod gas forces directly on suction and discharge pressure. By setting
F 1 greater than 1.0, we can account for the natural tendency of gas forces to exceed the values
based directly on suction and discharge pressure. Machines with good pulsation control and low
external resistance may achieve F1 as small as 1.1; for machines with low compression ratio, high
pulsations, or highly resistive piping external to the cylinder, F1 can approach 1.5 or even higher.

2-3
A reasonable working value for F1 is 1.15 or 1.2.

Preferably, the maximum rod force resulting from gas pressures would be based on knowledge of
the continuous variation of pressure in the cylinder (measured or predicted). In a repair situation,
measured cylinder pressure variation, for the compressor in question, obtained under the likely
range of operating conditions, using a cylinder analyzer, prior to analysis of the foundation design
would provide the most accurate value of gas forces. Even without cylinder pressure analysis
extreme operating values of suction and discharge pressure for each stage should be recorded prior
to the repair situation, and used in the above formulae.

On new compressors, the compressor supplier should be asked to provide a value for maximum
compressive and tensile gas loads on each cylinder, and to recommend a value for F1.

In the absence of such data from the OEM, particularly for a multi-stage compressor, the
Compressor Diagnostic Software (CDS; Ref. 1) provides one method to predict cylinder pressure
variation and calculate rod loads accounting for valve resistance. Alternatively, given knowledge
of interstage pressures, Equations 2-4 and 2-5 can provide values for maximum rod loads.

2.3 Compressor Inertia Forces

2 . 3 . 1 General

The following formulation develops expressions for inertia forces acting within the slider-crank
mechanism of an individual cylinder; such individual cylinder forces have direct relevance in
determining forces to be carried by compressor frame tie-downs.

2 . 3 . 2 Time Varying Inertia Forces

The motion of the piston, crank throws, and associated inertias involve time varying accelerations.
To accelerate the piston, piston rod, and crosshead requires time varying forces, mainly horizontal
(but with a small vertical component), reacted typically at the main and crosshead bearings.
Rotation of the crank throw, unless perfectly balanced, produces a rotating, unbalanced, force
vector, which in turn causes time varying horizontal and vertical forces, reacted at the main
bearings. The motion of the connecting rod involves a combination of rotating and reciprocating
motion, normally managed for predicting inertia forces by adding two-thirds of the connecting rod
mass to the rotating mass of the crank pin, and one-third of the connecting rod mass to the total

2-4
reciprocating inertia. If there are balance weights, they also each add a rotating force vector,
normally oriented to oppose rotating force vectors from the crank throw, and sometimes also to
partially offset the unbalanced reciprocating forces from the piston, etc. (alternative balancing
strategies is beyond the scope of the document).

The resultant inertia force variations are given as follows (from Ref. 5 and Appendix A):

Wri ri2ω 2
Fxi =
1
[ Wci + Wri ] ri ω cos (ωt + α i ) +
2
cos2 (ωt + α i ) (2-6)
g
g l

and

Wci
Fyi = ri ω 2 (ωt + α i ) (2-7)
g

where:

F xi = the horizontal inertia force required to accelerate the reciprocating and rotating
inertias, acting towards the main bearing, and causing a reaction on the main
bearing towards the cylinder (lb.).

F yi = the vertical inertia force required to accelerate the rotating inertias, causing an
upwards reaction on the main bearing (lb.).

αi = the phase angle for the crank radius of the ith cylinder, and measures by what angle
the ith crank leads the crank for the reference cylinder. For the reference cylinder,
the associated throw makes an angle of ωt to the cylinder centerline.

(These formulae apply the common approximation to account for connection rod angularity; this
creates two components of reciprocating force, one with frequency equal to rotational speed of the
shaft, and one at twice this frequency. There exist small higher order forces, but these normally
have no significance in foundation design.)

2-5
Note that the above formulae for horizontal and vertical forces do not account for balance weights;
Appendix A will provide a rigorous means to include balance weights in calculating summed
shaking forces and moments.

2 . 3 . 3 Maximum Inertia Forces

For horizontal compressor cylinders, we need the maximum value of unbalanced horizontal
shaking force for each cylinder, to assist with calculations of horizontal force to be restrained by
anchor bolt induced friction. The following provides a working approximation to this maximum,
including balance weight effects:

(Fx )i − max = rω 2 [(Wc + Wr ) + (1/n) Wr ] / g - ∑ (WR)bal- j ω 2 /g (2-8)

where (WR)bal-j represents the product of weight and eccentricity for the jth balance weight. The
balance weights are assumed to act in direct opposition to the line joining main bearing to crank pin

and

n = l/r (2-9)

where:

l = connecting rod length.


r = crank radius (stroke/2).

2.4 Summed Unbalanced Forces and Moments

2 . 4 . 1 General

The following forces and moments result from vector summations of inertia forces from individual
cylinders, accounting for phasing between crank throws. While widely used, these summed
forces should form the basis only for addressing the response of a complete foundation block upon
its soil support. Determining local forces at individual tie-down requires the use of the individual
cylinder forces developed above.

2-6
2 . 4 . 2 Eight Force and Moment Component Definition

In a multiple cylinder compressor, it is common to calculate eight components of unbalanced force


and moment amplitude:

• FH1, the horizontal primary force, with frequency of rotational speed.


• FV1, the vertical primary force, with frequency of rotational speed.
• FH2, the horizontal secondary force, with frequency of twice rotational speed.
• FV2, the vertical secondary force, with frequency of twice rotational speed.
• MH1, the primary moment acting in the horizontal plane.
• MV1, the primary moment acting in the vertical plane.
• MH2, the secondary moment acting in the horizontal plane.
• MV2, the secondary moment acting in the vertical plane.

2 . 4 . 3 Calculating Summed Forces and Moments

Appendix A provides a comprehensive calculation method for these forces and moments in a
multiple cylinder engine/compressor, with cylinders arrayed horizontally or at arbitrary angles to
the horizontal.

2.5 Separable Driver Torques and Forces

All compressors require some form of drive mechanism, either integral with the compressor or
separate from it; integral engine compressors incorporate both power cylinders and compressor
cylinders in the same frame; separable compressors incorporate only compressor cylinders in the
compressor frame and require a separate driver which applies a drive torque to the compressor
input shaft. Alternative drivers include:

• Electric Motor
• Reciprocating Engine
• Gas Turbine (normally through a gear box)
• Steam Turbine (normally through a gear box)

The drive torque of an integral engine/compressor balances the equal and opposite reaction between
driving and driven portion within the single engine/compressor frame, producing little or no net
external torque on the entire frame. However, the driver of a separable compressor produces a

2-7
distinct external torque on the compressor frame. The frames of both driver and driven machines
have equal and opposite torques acting on them, as shown in Figures 2-1a through 2-1c which
treat driver and driven as free bodies under torque.

For a separable compressor, the reaction torque applied to the foundation will typically appear as a
side-to-side difference in the vertical force acting to hold down the frame - both for driver and
driven machine. Figure 2-1d shows the mechanism, from which it may be inferred:

T = [ Fl - Fr ] w/2 (2-10)

and

F1 = F + 33,000 HP/(Nw)
Fr = F - 33,000 HP/(Nw)

where:

T = the external torque acting on the frame in question, with the frame viewed so that
this torque acts right handed on the frame.
F1 = the downward force acting on the left side of the frame as viewed.
Fr = the downward force acting on the right side of the frame as viewed.
w = the transverse separation between the lines of action of Fl and Fr.
HP = the horsepower.
N = the rotational speed in RPM.

And it is assumed that the average of the forces acting on either side of the frame ( F ) is
unchanged.

2.6 Managing Inertia Forces and Moments

2 . 6 . 1 Global Considerations Based on Summed Forces and Moments

Summed unbalanced shaking forces and moments for the compressor or engine/compressor act on
the entire structure (machine and foundation). The structure rests on flexible soil, giving strong
reason to manage these forces and moments, and to limit dynamic response of the structure on the

2-8
soil. Managing forces and moments depends in part on internal balancing. Changing phase angle
between throws of a multi-throw crankshaft, studied selection of added balance weights, and
accurately matching nominally equal weights enable control of the eight unbalanced forces and
moments; in many cases, most of the summed forces and moments can be made small or zero.
However, with eight components in total, some compromise is often required and some forces or
moments may remain high. It is particularly important to avoid combining a high shaking force or
moment with a machine-soil natural frequency whose mode shape and frequency will respond
sympathetically to the form and frequency of the high force or moment. This combination can lead
to significant global motion of the compressor with possible damage to piping and buildings, and
misalignment of cylinders.

Thus, dynamic considerations of the summed shaking forces and moments, combined with natural
frequencies of the structure on the soil must receive careful attention. These guidelines do not
address the details of dynamic response of the foundation on its soil mount, but a number of texts
do (Refs. 10, 11, and 12).

2 . 6 . 2 Local Considerations

Note that summed forces and moments act on the entire structure, and may bear little relationship to
forces acting at interfaces within the structure. Adding counterweights on a throw can locally
reduce the bearing force, but balancing two large forces in adjacent cylinders by making them act
out of phase on the entire structure can still leave large forces acting on the bearings and other
internal components of the structure. In particular, this warning manifests itself in forces
transmitted to individual anchor bolts, and through them to subsections of the foundation. These
local forces can overload individual tie-downs and overstress locally vulnerable areas within the
foundation. Managing global and local forces in a compressor foundation structure requires
knowledge of the individual forces acting within a cylinder and throw.

2.7 Managing Local Gas Forces

As discussed further below, the distinction of local and global balancing applies to gas forces as
well as to inertia forces. Gas forces are reacted at main bearings and at the joint between cylinder
and crosshead guide. For the compressor/foundation structure as a whole, the two gas force
reactions internally balance, naturally, but along the structural path through which this internal
balancing is accomplished, there may be large local forces. The fact that cylinder supports are kept

2-9
separate from the main foundation recognizes a phenomenon called cylinder stretch, but in fact,
frame and crosshead guide stretch also occur and can put high loads on crosshead guide tie-downs
and on the foundation to which they are attached.

Local gas forces must be considered in engineering tie-downs, and in controlling foundation block
stresses.

2-10
Driver Driven

FIGURE 2-1a. PLAN VIEW OF DRIVER CONNECTED TO DRIVEN MACHINE

FIGURE 2-1b. FREE BODY DIAGRAM IN PLAN VIEW, SHOWING TORQUES


FROM ATTACHED MACHINE AND REACTION TORQUES FROM
FOUNDATION

2-11
Driver Driven

+ +

FIGURE 2-1c. FREE BODY DIAGRAM IN END ELEVATION LOOKING FROM


DRIVER END TOWARDS DRIVEN MACHINE SHOWING TORQUES FROM
ATTACHED MACHINE, REACTION TORQUES FROM FOUNDATION, AND
TORQUES APPLIED TO FOUNDATION

FL FR
• T

FIGURE 2-1d. FREE BODY DIAGRAM FROM A FRAME IN WHICH DRIVE


TORQUE IS REACTED BY DIFFERENT NET FORCES ON EITHER SIDE OF
THE FRAME

2-12
3.0 TRANSMISSION OF FORCES

3.1 Overview and Structural Considerations

This section addresses the transmission of forces from origination point to reaction points
(bearings, compressor frame, compressor mounting, foundation, and soil). It distinguishes gas
and inertia forces. It emphasizes that all forces transmit themselves through a structure by
deforming that structure. The structure subjected to gas and inertia forces is comprised of
compressor, concrete foundation, and the interface between them. Both concrete and the metal
frame are flexible, and transmission of forces deform both concrete and metal frame. We have
interest in determining forces and stresses at various locations, in particular, where the compressor
connects to concrete and at high stress points in the concrete. Determining such forces requires
careful attention to the deformation process and awareness of the relative flexibility of frame and
concrete under different types of loads.

In general, the following sections will show that frame flexibility increases forces transmitted to the
block via the mounting system.

3.2 Gas Force Transmission

Gas forces act on the crankshaft with an equal and opposite reaction on the cylinder. Thus,
crankshaft and cylinder forces globally balance each other; however, the internal balancing
mechanisms deserve careful review. Between crankshaft and cylinder, the compressor structure
stretches or contracts in tension or compression under equal and opposite gas forces. The
connections between compressor frame, crosshead, and the concrete transmit frame deflections to
the foundation block. In fact, when acting without slippage (as designed), the frame and block
become an integral structure and together stretch or contract under the gas loads. The sharing of
each cylinder’s local gas force between compressor and foundation depends on relative flexibility
of the two structures. A highly rigid compressor frame would transmit only a small fraction of the
gas force; a very flexible frame would transmit most or all of the gas forces to the block. The
normally “moderate” flexibility of both frame and block causes load sharing somewhere in between
these extremes. In subsequent sections results of finite element analysis will illustrate how frame
and block flexibility influence transmitted loads.

3-1
3.3 Inertia Force Transmission

3 . 3 . 1 How Flexibility Impacts Local Forces

Substantial, time-varying inertia forces act on each crank throw and its bearings but not on the
cylinder. This “unbalance” distinguishes inertia from gas forces. As previously discussed, the
phase relationship between throws directly influences the net inertia forces and moments applied to
all the bearings as a whole.

However, even with summed forces and moments, small or zero, compressor frame bending
flexibility ensures tie-downs and foundation near a crank throw experience a substantial fraction of
that throw’s shaking forces. With an extremely rigid frame, minimizing the summed forces and
moments should also minimize the locally transmitted forces. A very flexible frame makes locally
transmitted forces essentially independent of the summed forces and moments. Reality lies
between the rigid and very flexible extremes.

3 . 2 . 2 Measured Flexibility

Figure 3-1 illustrates compressor frame flexibility through a display of measured frame vibration
at a series of crankshaft rotation angles (every 45°); clearly, the frame motion includes substantial
bending, as well as rigid body motion.

3.4 Calculation Methods

Various standard methods exist for calculating the forces transmitted to the foundation block and to
each tie-down which connects the frame to the block. The major difference between these methods
lies in their treatment of frame and CHG flexibility, and in the effort they require. Different
assumptions regarding these flexibilities, summarized in the following subsections, can
significantly influence the calculated values for transmitted forces.

3 . 4 . 1 Conservative Calculation Method

Assume a totally flexible frame and crosshead guide. This means the tie-down and foundation
strength, local to a cylinder, must carry the full maximum gas force in that cylinder, and must carry
inertia forces associated with that crank throw. While relatively simple to implement, given the gas
and inertia forces, this approach can overestimate forces, but is preferred to 3.4.2.

3-2
3 . 4 . 2 Non-Conservative Calculation Method

Assume a rigid frame. This assumption allows sharing of net unbalanced forces and moments
between all anchor bolts and neglects gas forces. While simple to implement, it underestimates the
transmitted load and anchor forces - often by a very substantial amount, and is not recommended.

3 . 4 . 3 Finite Element Method

Construct and apply a detailed model which accounts effectively for the individual contributions of
compressor frame flexibility and foundation flexibility to the total transmitted force. While a more
complex approach, this produces forces more realistic than either simplifying assumption.
Appendix B describes this approach. Subsequent sections of this document show how the
different simplifying assumptions compare with this more rigorous approach.

3 . 4 . 4 Guidance from Previous Rigorous Analysis

3.4.4.1 Basis for Guidance - Appendix B summarizes a detailed analysis (Ref. 9)


performed on a six cylinder horizontally opposed compressor mounted on a
massive concrete block foundation. Modeling the compressor frame and
foundation as a single structure provides values for forces at the interface between
frame and block. These results can act as a reference for evaluating the same
interface forces obtained using simplifying assumptions. A number of guidelines
(see 3.4.4.2, 3.4.4.3, 3.4.4.4, 3.4.4.5, and 3.4.4.6) result from this study.

Figure 3-2 presents in bar graph format a comparison of forces obtained using the
three calculation methods, extracted from Appendix B. Close study of this figure
makes clear the gross underestimate obtained from the non-conservative method;
the moderate over-estimate obtained with the conservative method; and the rationale
for a modest reduction of forces obtained with the conservative method. The
following guidelines result directly from study of Figure 3-2.

3.4.4.2 Calculating Horizontal Forces to be Restrained (General) - Use the


forces associated with individual cylinders as a basis for calculating horizontal forces to be
restrained at anchor bolts.

3-3
Do not use summed forces or moments for the complete compressor as a basis for
calculating horizontal forces to be restrained at anchor bolts. Although using
summed forces and moments is quite common practice in foundation engineering
for reciprocating compressors, the required restraining capacity for individual bolts
obtained in this way bears little or no relationship to reality, and can grossly
underestimate this required tie-down capacity by more than 90 percent.

3.4.4.3 Forces Transmitted by Crosshead Guide - Use individual cylinder gas


forces (maximum tensile rod load) to obtain a horizontal stretching force acting on
the top of the block near the crosshead guide for calculating concrete stresses which
result from applied dynamic loads.

3.4.4.4 Required Tie-Down Capacity at CHG Bolts - Use individual cylinder gas
forces to analyze required tie-down capacity at crosshead guide anchor bolts.
Divide maximum rod load from gas forces by the number of crosshead guide bolts
to obtain required holding capacity.

3.4.4.5 Required Tie-Down Capacity at Frame Bolts - Use individual cylinder


inertia loads as a basis for analyzing the required tie-down holding capacity for
anchor bolts along the sides of the frame. Divide the maximum inertia force per
throw by the number of frame bolts on both sides of the frame, which can be
reasonably assigned to that throw. A simple way to make this calculation is to
multiply the maximum throw unbalance force by the number of throws, then divide
by the total number of anchor bolts arranged along the sides of the frame (excluding
crosshead guide bolts).

3.4.4.6 Safe Force Reduction Factors - Based on the limited experience of one study
on a relatively flexible compressor frame, it is probably safe to reduce by a factor of
two all the forces obtained using 3.4.4.4, 3.4.4.5, and 3.4.4.6. This preserves the
local gas and inertia forces as the drivers of loads on tie-downs and compressor
block which influence integrity, but acknowledges the over-conservatism of the
simple treatment of Sections 3.4.4.4, 3.4.4.5, and 3.4.4.6.

3-4
3.4.4.7 Summary of Force Formulae:

Fblock = (Frod) gt-max/Fred (3-1)


(Fbolt)CHG = (Frod) gt-max/(Nbolt)CHG/Fred (3-2)
(Fbolt)frame = (Fx)i-max/(Nbolt)frame/Fred (3-3)
with Fred = 2.0 (3-4)

where:

F block = the force acting outwards on the block from which concrete stresses should
be calculated.

(Fbolt)CHG = the force to be restrained by friction at the crosshead guide tie-down bolts.

(Fbolt)frame = the force to be restrained by friction at the frame tie-down bolts.

and

(Frod) gt-max comes from Equation 2-5, and (Fx)i-max comes from Equation 2-8.

3-5
FIGURE 3-1. DEFLECTED SHAPES OF A COMPRESSOR FRAME FOR EACH
45 DEGREES OF CRANKSHAFT ROTATION

120000

100000

80000
Soft Frame
60000 "FE"
Rigid Frame
40000

20000

0
Block CHG FRM
Force Bolt Bolt
Force Force

FIGURE 3-2. INFLUENCE OF ASSUMPTIONS ON FORCES TO BLOCK, CHG


BOLTS, AND FRAME BOLTS

3-6
4.0 COMPRESSOR MOUNTING

4.1 Overview and Requirements

Compressor mounts must robustly accommodate all loads applied by the frame - horizontal and
vertical gas and inertia loads, weight loads, vertical anchor bolt loads applied through the base of
the frame, and loads resulting from thermal growth and distortion of frame and block. All but
horizontal thermal loads should be directly transmitted to the concrete block by the mount. The
exception implied for thermal loads will be discussed in more detail below.

4.2 The Role of Friction

4 . 2 . 1 Relating Tangential and Normal Forces

Friction provides the primary mechanism for transferring dynamic horizontal forces from the base
of the compressor frame to the mount. To avoid slippage under dynamic loads, the normal force at
the interface multiplied by the coefficient of friction must exceed the maximum horizontal dynamic
force applied by the frame at the tie-down. To assess adequacy of a compressor tie-down requires
knowledge of the horizontal force the tie-down must restrain, the coefficient of friction, the tensile
capacity of the anchor bolt and the tensile capacity of the anchorage of the bolt. Put another way,
knowledge of the friction coefficient and the horizontal load to be restrained sets the minimum
tensile preload in the bolt to assure restraining the horizontal load:

Tmin = Fr/µ - Wr (4-1)

4 . 2 . 2 Anchor Bolt Requirement

An anchor bolt, which with its concrete anchorage, has long term tensile strength in excess of Tmin,
maintained at or above this tension, coupled with a chock interface whose coefficient of friction
equals or exceeds µ, will hold the force Fr to be restrained. Here, Wr, represents the share of the
compressor's weight which acts at this tie down. A conservative approach would neglect Wr (set it
to zero) since vertical distortion of frame or block may reduce the effective weight force at this
location below Wr.

For simplicity, the largest force to be restrained for frame or crosshead guide should be used as Fr
for design of all the anchor bolts according to Equation 4-1.

4-1
4 . 2 . 3 Friction Data

The PCRC research program has set out to develop data for friction between common chock
interface materials, including steel/cast iron, steel/steel, epoxy/cast iron, epoxy/steel. epoxy/epoxy,
dry and with oil present, using sizes which come close to those typical of compressor mounting
practice. This data, when available, will supplement Equation 4-1. Reference 14 provides some
data for steel and cast iron.

4.3 Chock Configuration

The bolt holes in the compressor frame and in the crosshead guide fix the tie-down locations (and
anchor bolt diameter). Figures 4-1 and 4-2 show some of the chock configurations which can be
considered. The first two schematics of Figure 4-1 include poured in-place chocks, which have
particular advantage in repair situations, where the compressor base may have imperfections due to
past wear, fretting, or corrosion. The second, third, and fourth schematics of Figure 4-1 include
adjustable chocks which use shims, or a shallow tapered interface (which provides vertical
adjustment by inducing tangential motion between the two tapered surfaces). Both types can be of
steel composite material or epoxy. Also shown in Figure 4-1 is the full bed grout, rarely used in
new installations or repairs today, except in a few high pressure compressor installations for
polyethylene; however, many full bed installations exist, still running from original installation
twenty-five or more years ago. The lips shown in the first two installations occur as a natural
result of the pouring process; however, they do not have structural value and tend to break off
when exposed to relative thermal expansion between frame and block. Some installation practices
will specifically remove these tips after pouring. Figures 4-2 and 4-3 present photographs of
several installations using different chock or mount configurations, including full bed, adjustable
epoxy, and adjustable steel.

4 . 3 . 1 Benefits of Chock Mounting

Benefits of chock mounting include:

• An extensive air gap, which inhibits heat flow from frame to block, and thereby reduces block
distortion.
• A series of positive tie-down points with known, engineerable, interface conditions.
• Adjustability.

4-2
The choice of epoxy versus steel remains a controversial subject; both have their advantages and
their disadvantages (and their advocates).

4 . 3 . 2 Steel's Advantages

Steel is very strong; steel is stiff; steel does not creep; steel chocks can be made as adjustable as
epoxy chocks by shimming.

4 . 3 . 3 Steel's Disadvantages

Steel cannot be poured in place; steel has a higher conductivity than epoxy.

4 . 3 . 4 Epoxy's Advantages

Epoxy is pourable; epoxy has lower heat conductivity that steel; epoxy can have higher friction
coefficient than steel; epoxy can be made adequately strong.

4 . 3 . 5 Epoxy's Disadvantages

Epoxy creeps; epoxy is less stiff than steel; epoxy is less strong than steel; some epoxies are brittle.
Epoxy has a higher coefficient of expansion than steel, cast iron, or concrete.

4 . 3 . 6 Chock Design

In general, given knowledge of each chock’s material properties, an adequate chock using either
material can be designed for most if not all configurations. The problem of creep in epoxy implies
a need for periodic tightening of the anchor bolts to accommodate the reduction in preload as the
compressed area of epoxy continues to deflect with time. There is a need for additional data and
characterization of epoxy's creep behavior, to support definition and documentation of installation
and maintenance requirements, and frequency of adjustment requirements.

4 . 3 . 7 Normal Load Limits

A chock is subject to compression loading by the combination of compressor weight and anchor
bolt tension. These act on the chock surface area to produce a net pressure load, which should be
below the allowable pressure for the chock in question:

4-3
where Pchock = (T + Wr)/( l c * wc) < Pallow

The chock supplier should be able to provide a recommended or allowable pressure load (typically
500 to 1000 psi for epoxy, and substantially higher for steel). Compressive strength for epoxies
typically exceeds 10,000 psi, and substantially higher for steel.

where:

P chock = applied pressure loading.


T = anchor bolt tension.
Wr = portion of compressor and engine weight assignable to this chock.
wc = chock width.
lc = chock length.
Pallow = allowable chock pressure loading.

If the required anchor bolt load gives rise to a chock pressure loading which exceeds allowable,
then the options available are:

• Increase chock surface area.


• Increase chock allowable load by material change.
• Reduce the required anchor bolt load.

An engineering evaluation of these alternatives will lead to one or more solution options, and a
mount of satisfactory strength.

4 . 3 . 8 Effects of Creep and Anchor Bolt Length

James and Winston (Ref. 15) provide data on creep for epoxy grout materials, with reference to
reciprocating compressor mounts. They describe creep for epoxy materials as time dependent
deflection under load and present graphs of creep compliance as a function of time and
temperature. Creep compliance quantifies strain per unit stress and is essentially the inverse of
Young’s Modulus. This representation implies all epoxy materials creep, that the deflection for a
given time is proportional to load, and that deflection increases with time and temperature.

4-4
Typical values presented in this publication for creep compliance range from 0.7 x 10-6 to 10 x 10-6
in2/lb. One material shows a relatively rapid change in deflection with time, and another appears to
be soft but relatively stable with time. Data for three of the five materials tested suggest that a
mount can readily be designed which in 10 years will change deflection, under 1,000 psi chock
loading at 130°F, by less than 6 mils for a 1-inch thick chock (much less for some materials). In a
30-inch long bolt, the resultant “worst case” loss in stretched length for this creep rate would be
200 microstrains, corresponding to a loss in tensile stress of about 6,000 psi - 15 percent for a bolt
tensioned to 40,000 psi. Increasing the bolt length to 60 inches would reduce this loss in tension
to 3,000 psi (7.5 percent), and for a 10 foot bolt, the loss in tensile stress would fall to only 1,500
psi (3.8 percent). Some epoxies will creep much less than 6 mils, so epoxy creep appears to be
manageable once knowledge of the phenomenon is available and applied.

In summary, understanding the creep process for epoxy grout materials and availability of creep
data should enable the mounts and anchor bolt tension to be engineered for desired applications.

4.4 Anchor Bolt Preload

The anchor bolt preload must provide sufficient normal force across horizontal interfaces between
base of frame, chock, and soleplate, so that the maximum dynamic horizontal force from the
compressor to be carried at the interface does not exceed the frictional holding capacity (normal
force multiplied by coefficient of friction), as defined in Equation 4-1. Since holes in the frame
and crosshead guide set bolt diameter, the bolt material’s yield limit sets the maximum possible
preload. However, high strength anchor bolts of material conforming to ASTM A-193 (Ref. 16)
now offer a tensile yield strength of 105,000 psi.

Example:

A gas load of 108,000 lb. is to be carried by four crosshead guide bolts of 2 inches in nominal
diameter.

The nominal gas load per bolt is 27,000 lb.

Load sharing by frame deflection is expected to reduce the required per bolt horizontal load by a
factor of 2, based on guidelines of Section 3.4.4.6.

4-5
For a coefficient of friction of 0.15, the required preload is:

27,000/ (2 * 0.15) = 90,000 lb.

4.5 Thermal Distortion and Growth

Sump oil temperature typically ranges from 140°F to 160°F and maintains the bottom of the frame
at approximately this temperature. Heat from the bottom of the frame flows to cooler regions by
paths of least resistance. In particular, the heat flows to the foundation block through the chocks
(and to a lesser extent through the air gap). Thus, the top of the block tends to get hotter than the
bottom of the block by 30°F or more.

4 . 5 . 1 Block Distortio n

Considering the block as a beam, the temperature gradient from top to bottom will tend to cause the
block to bow up.

4 . 5 . 2 Frame Distortion

It is quite widely believed such block distortion will impose itself on the compressor frame. In
some cases it does. However, it is not assured that the frame at the crankshaft level will
automatically, or consistently assume this anticipated deflected shape of the foundation block.
Evidence exists from more than one source that the combined response of the compressor and
block structure to imposed temperature distribution can cause the frame to bow up or down, with
downward bowing particularly likely during the first 24 hours of operation.

4 . 5 . 3 Cyclic Frame Distortion

Available data further suggests that daily and seasonal temperature cycles will cause the amount of
bow observed at the bearing line of a compressor to cycle up and down. Specific data (Ref. 8)
exists for a 20 foot long HBA-8 compressor, which exhibited cyclic variation in alignment of
center relative to the ends with 12 thousandths of an inch range over a four day period in summer;
wider variation could be expected over a larger period of time with wider temperature swings.

4-6
4 . 5 . 4 Alignment Implications of Thermal Distortion

The uncertainty discussed in Sections 4.5.2 and 4.5.3 suggests aligning the compressor to level
versus attempting to compensate for upward bow which may not always occur.

4 . 5 . 5 Weight Load Implications of Thermal Distortion

Whatever direction thermal distortion can occur, it will likely cause increased or decreased normal
force at individual frame to chock interfaces. This suggests neglecting the weight of the
compressor in anchor bolt tension calculations; however, in analyzing the influence of normal load
relative to strength of the chock, including the weight, represents a conservative approach.

4 . 5 . 6 Horizontal Load Implications of Thermal Growth

Even in the absence of thermal bowing, differential axial and lateral growth between frame and
block must occur. Observations suggest that resistance to heat transfer can cause as much as 30°F
temperature differential between bottom of frame and top of block. The differential growth in the
axial dimension of a 25 foot long compressor frame and block can exceed 50 mils, and the
differential lateral growth between outboard crosshead guide mounts 12 feet apart can exceed 25
mils. If shared without slipping between the outermost tie-downs, such differential growths
across chocks 1 to 2 inches thick would cause shear strains between 6,000 and 25,000 microstrain
and require a shearing force of a million pounds or more. Thus, slipping under static thermal
growth appears inevitable in many of the epoxy chocks, particularly those at the corners. In
anticipation of this requirement to slip the majority of poured epoxy chock providers use a mold
release agent to help ensure that bonding does not occur at the frame to chock interface. With steel
chocks such high shear strains also ensure that slipping will occur.

To summarize the points made in subsection 4.5.6:

• Install and align compressors level.


• Do not take credit for compressor weight in calculating anchor bolt tension to hold a particular
horizontal dynamic force with friction.
• Do account for weight in chock strength calculations.
• Anticipate sliding at chock mounts to accommodate differential thermal growth.

4-7
• Use a release agent to avoid adhesive bonding in poured epoxy chocks, unless specific
application data indicates otherwise.

4-8
FIGURE 4-1. TYPES OF COMPRESSOR FRAME SUPPORT SYSTEMS

(ILLUSTRATIONS COURTESY OF ROBT. L. ROWAN & ASSOC., INC.)

4-9
SHIMMABLE COMPOSITE CHOCKS ON STEEL SOLEPLATES

SHIMMABLE STEEL CHOCKS ON STEEL SOLEPLATES

FIGURE 4-2. ADJUSTABLE CHOCKS

4-10
FIGURE 4-3. FULL BED GROUT

4-11
5.0 ANCHOR BOLTS

5.1 Material

5 . 1 . 1 High Strength Material for New Installations

Anchor bolts need high strength to provide design and operational flexibility to meet demanding
requirements to hold against high dynamic forces. All new installations should use high strength
steel anchor bolts satisfying ASTM specification A-193 (Ref. 16), with 105,000 psi tensile yield.
For a given diameter, this maximizes the capacity of the anchor bolt to provide required preload.

5 . 1 . 2 Lower Strength Bolts in Old Installations

Most older installations have anchor bolts with much lower tensile yield strength (e.g., 40,000
psi). For foundation repairs which retain part of the old anchor bolts, techniques exist for coupling
an existing bolt section of low strength extending into old concrete to a new high strength bolt
section extending from the coupling to the compressor frame, and supplementing the limited
strength of the old bolt section. This subject will be further addressed under the subheading length
and configuration.

5.2 Anchor Bolt Diameter

The compressor frame configuration normally sets the number and diameter of anchor bolts,
limiting the options available for maximizing preload capacity to choice of material, length, and
thread configuration. This illustrates clearly the system aspects of compressor installation, the
overlapping of responsibilities between organizations, the importance of this particular design
decision for the compressor frame, and its ability to influence the integrity of the entire installation.
Thus, compressor manufacturers should carefully review their decisions on frame holes in light of
their impact on anchor bolt preload capacity and tie-down holding capacity.

5.3 Threads

Use rolled (formed) threads in all new anchor bolt specifications to minimize any stress
concentration at the threads (as per ASTM A-193 for bolts and ASTM A-194 for nuts).

5-1
5.4 Length

Make anchor bolts as long as possible, for the following reasons:

• Lengthening anchor bolts increases their tolerance to creep of any polymeric materials in the
machine mounting. Even with steel chocks, current installation practice for new or repaired
foundations always includes epoxy grout, so the potential for creep always exists. With
creep - sensitive components (chocks, grout) localized near the top of the foundation block,
doubling the anchor bolt length cuts in half the loss of bolt preload, which will occur as a result
of polymer creep.

• Lengthening the anchor bolts reduces the dynamic stresses in the concrete near the bolt
termination. Horizontal and vertical dynamic loads act at the top of the foundation and cause
high localized time varying stresses near the point of application. Well down in the foundation,
such dynamic stresses are minimized. The anchor bolt termination can cause high tensile stress
and cracking. Because cracks subject to dynamic loads tend to grow, moving the termination
point far from the dynamic stress increases the chance of long term integrity.

• Lengthening anchor bolts increases the separation of bolt termination from sources of oil. The
presence of oil at any point in a crack leads to migration of oil along the crack. Oil can help
grow the crack and weaken the surrounding concrete. Since cracking at the termination is hard
to avoid, maximizing the separation of oil sources and the bolt termination helps ensure long
term integrity of the installation.

• Carried to the ultimate, maximizing anchor bolt length puts the termination in the mat on which
the foundation is mounted, rather than in the foundation block itself. There exists a growing
practice to put the anchor bolt termination on new and even repaired installations below the
horizontal mid-plane of the mat. Beyond length maximization, this means that any cracks
induced by tensile stresses at the termination cannot grow outwards to the side of the block,
where they are unsightly, cause concern, and may reduce integrity of the installation. In old
installations penetrating the mat may encounter old rebar, but the overall improvements in
integrity should offset any such local loss of rebar. Practitioners of this approach to anchor
bolt installation claim it is the easier than terminating the bolt in the poured concrete of the block
itself.

5-2
5.5 Configuration

In new installations a long bolt terminating with a small circular plate or simply a nut provides the
best configuration. In old installations a variety of configurations may exist. Figure 5-1 shows
some alternatives.

5 . 5 . 1 Avoid J and L B o lt s

The first two configurations, J and L bolts, while widely used in the past, should never be
considered in new installations. Their asymmetry, inability to develop a symmetrical cone of
compression, and predicted high potential for cracks around the termination make them
undesirable. As evidenced by tests supported by PCRC, if overloaded, such bolts start to yield
near the termination. They continue to fail by straightening out the bend. Having no other
diameter increase device ( such as a nut or plate), the anchor bolt can then pull up through the hole
in the surrounding concrete.

5 . 5 . 2 Symmetrical Terminations

All other terminations in Figure 5-1 are symmetrical and can develop a symmetrical cone of
compression in the concrete under bolt tension. This is desirable.

5 . 5 . 3 Square or Rectangular Terminations

Use of a square or rectangular termination is undesirable because it lacks symmetry. Analysis


shows a tendency to form local high tensile stress and cracks at the steel to concrete interface near
the corners.

5 . 5 . 4 Large Termination Plates

Use of large termination plates does not increase integrity and can extend the area over which
termination induced cracking can occur. If a plate termination is used, such as the third
configuration in Figure 5-1, it should be circular and small.

5-3
5 . 5 . 5 Nut Terminations

The nuts used in the fourth and fifth configurations of Figure 5-1 provide fully adequate
termination for high strength anchor bolts. They bear on the concrete above them and ensure
developing a cone of compression which will adequately anchor the bolt in the concrete.

5 . 5 . 6 Through to the Mat Configuration

The fifth configuration in Figure 5-1 represents the through to the mat configuration discussed as
the ultimate and optimum way to maximize anchor bolt length. As shown, the bolt extends more
than halfway into the mat, terminating in a nut. This configuration nearly triples the length of bolt
relative to the J, L, and plate bolt configurations, with the aforementioned benefits in long term
integrity.

5 . 5 . 7 Canister Bolt

The sixth configuration in Figure 5-1, entitled “canister bolt,” has value in repair situations where
the anchor bolts must be replaced, but the concrete will not be removed to the point where the bolt
should terminate. A core drilled hole, wide enough to accommodate the canister enables this
configuration to be installed and epoxied in place. The bolt, attached to the canister at the bottom,
uses the canister as a sleeve, and develops tension by stretching this free length between the point
of connection to the canister and the compressor frame. This configuration can serve as an anchor
bolt, or as a supplementary post-tensioning bolt which can place areas of concrete, such as a
cracked region in compression and add integrity. Its design enables the top of the bolt to be
temporarily lowered to the top of the block.

5 . 5 . 8 Supplementary Strength of Old Anchor Bolts

A previous subsection discussed an approach which can achieve anchor bolt strength approaching
that of a high strength steel, even if it involves coupling to an old lower strength bolt - perhaps
with a J or L termination.

Figure 5-2a shows a finite element model of a foundation with an old L bolt to which a coupling
with supplementary tendons has been attached. Finite element results show that addition of the
tendons can reduce the tendency for cracking of the concrete around the L termination. While a full
replacement with longer anchor bolts would be a desirable alternative, use of such techniques in

5-4
partial repair can optimize the use of constrained repair budgets and achieve integrity which
approaches that of a full high strength bolt installation. Figure 5-2b shows one supplier’s
installation schematic for such a bolt extension.

5.6 Preload

The desired preload is that load necessary to achieve adequate frictional holding capacity as defined
in Equation 4-2. The suggested allowable preload (for high strength steels) is that which does not
cause a bolt stress higher than 70 percent of tensile yield strength. For low strength steels, stresses
as high as 90 percent of yield are tolerable.

5.7 Sleeving

5 . 7 . 1 Relief Near Top of Bolt

It is undesirable and unrealistic for the bolt to maintain intimate contact with the concrete over its
entire length. Near the top there may be a need to move the bolt around during installation of the
compressor so the bolt can pass through the bolt holes in the compressor frame. In addition,
relative thermal growth of the frame may create a need for some freedom to move in the horizontal
plane. If rigidly held by the concrete too near the top, such adjustments or movements could
deform the bolt or introduce undesirable cracks in the concrete. Recommended practice is to leave
an air gap or soft sleeve between the bolt and concrete from the top of the bolt down ten bolt
diameters into the concrete, so that any needed bending can be accommodated without high stress
and without imposing high forces on the concrete where it is in contact with the bolt.

5 . 7 . 2 Bolt Stretch Sleeving

Since the bolt must stretch to achieve preload, relative lengthwise motion between bolt and concrete
will occur, reaching a maximum near the top of the bolt. A ten foot bolt, subject to 70,000 psi
tensile stress will stretch over a quarter of an inch. Such relative motion forces local breakdown of
any bond between concrete and bolt surface. In anticipation of this relative motion and to avoid
damaging the concrete, it is desirable to enforce separation between the bolt and the concrete over
the length of the bolt. This is accomplished using tape wrapped around the bolt or a thin-walled
polyethylene pipe around the bolt at installation. With the latter approach it is important that the OD
of the polyethylene pipe be significantly less than the diameter of the termination restraint.

5-5
The canister bolt illustrated under the Configuration section is one approach to sleeving and clearly
allows the desired unencumbered axial growth of the bolt relative to the concrete.

5.8 Nuts and Washers

The interface between anchor bolt and the frame deserves careful attention as does every other
detail of the installation. The specific configuration of nut and washer can influence the integrity of
the installation.

5 . 8 . 1 Spherical Self-Aligning Washers

Relative axial motion between frame and block during installation and operation must cause some
small change in angularity between anchor bolt and frame surface. Imperfections and tolerances
inevitably cause some lack of squareness. A self-aligning washer provides benefits at this location
and does not appear to have any disadvantages. A common configuration involves two pieces with
a spherical interface, as illustrated at the top of the canister bolt in Figure 5-1. The sphere can be
convex up or down.

5 . 8 . 2 Hex N u t s

Conventional hex nuts are widely used in compressor installations. ASTM Specification A-194
addresses nuts for high strength bolts. They require care to achieve the needed tension through
torque measurement. Investigations have shown that wide variability in tension can occur with
nuts tightened to the same torque. Lubrication of the threads helps ensure uniformity. Use of one
of the tension monitoring bolts on the market can also help overcome this problem. Ultrasonic bolt
stretch monitoring can also provide an independent check on preload. Hydraulic tensioners can
also assist.

5 . 8 . 3 Super N u t s

A device which helps achieve uniformity of tensioning is the Multi Jackbolt Tensioner (MJT),
illustrated in Figure 5-3. Instead of a single hex nut, a series of jackbolts tightened with Allen
wrenches to a relatively low and controllable torque, can provide the same tension as a much more
highly torqued hex nut. This patented device is increasing in popularity. It requires careful

5-6
sequencing in the processes of both tightening and loosening to avoid overload and difficulty of
further adjustment. In one example, 15 ft./lb. on each jackbolt produced the equivalent of 700
ft./lb. for a 1.25 inch bolt!

5.9 Anchorage

5 . 9 . 1 Strength Experiences

In all analyses performed, in spite of localized cracking near the bolt termination, there has been no
prediction of an anchor bolt crushing the concrete above it and pulling out. Appendix C describes
some pull-out tests. Some recent experience can be summarized as follows:

• A 2 inch diameter high strength anchor bolt with 120,000 lb. specified preload has been
installed without difficulty and has maintained its tension in 4,000 psi (compressive strength)
concrete. The termination was a 6 inch circular plate held in place with a nut. Some of these
bolts required torquing to over 3,000 ft.-lb. on the hex nut to put the tension monitoring bolt
into the tolerance range for 120,000 lb. tension.

• In two pullout tests on J bolts, the first observation was yielding of the bolt near the
termination followed by the anchor bolt pulling out by straightening of the J (see Appendix C).

• In a pullout test on a higher strength symmetrically terminated bolt, the first indication of failure
was a classical yielding characteristic of the bolt material (see Appendix C).

5 . 9 . 2 Design of Anchorage

The above results suggest that with a symmetrical termination securely attached to the anchor bolt
in concrete with 4,000 psi compression strength or more, the mode of failure will be yielding
before pulling out. Thus, designing the anchor bolt for the preload should ensure a secure
anchorage.

5-7
J BOLT

L BOLT

PLATE BOLT

2 PIECE ALLOY

THROUGH TO THE
MAT

CANISTER
BOLT

FIGURE 5-1. EVOLUTION OF ANCHOR BOLT DESIGNS


(ILLUSTRATION COURTESY OF ROBT. L. ROWAN & ASSOC., INC.)

5-8
FIGURE 5-2a. MULTIPLE TERMINATION POINT
ANCHOR BOLT SUPPLEMENT

5-9
FIGURE 5-2b. MULTIPLE EXTENSION METHOD OF INCREASING
ANCHORAGE STRENGTH FOR OLD ANCHOR BOLTS
(ILLUSTRATION COURTESY OF ROBT. L. ROWAN & ASSOC., INC.)

5-10
FIGURE 5-3. MULTI JACKBOLT MECHANICAL TENSIONER
INSTALLATION ON CROSSHEAD GUIDE TIE-DOWN

5-11
6.0 FOUNDATION BLOCK

6.1 Overview

The foundation block has the role of positioning and supporting the compressor, providing a fixed
reference upon which to maintain the compressor’s alignment, and acting as an extension of the
compressor’s structure in controlling motion of the compressor under dynamic loads. This latter
role involves adding both mass and stiffness. In both resonant and non-resonant situations, the
magnitude of vibration response to dynamic loads varies inversely with the vibrating mass. Thus,
adding significant mass is good and widely applied practice is to design a foundation with two to
three times the weight of the compressor itself. The role of adding stiffness to the structure has
been discussed in previous sections relative to the transmission of forces.

Achieving this role involves satisfactory attachment of the compressor to the foundation, a subject
extensively covered in preceding sections. When followed, the preceding guidelines should ensure
an intimately connected compressor and foundation, with direct transmittal of all compressor loads
to the foundation. Thus for full integrity, the concrete block foundation must tolerate the applied
loads.

6.2 Concrete

6 . 2 . 1 Positive Characteristics for Compressor Foundations

Concrete, a composite material with several valuable properties, has some characteristics that create
a challenge for designers of machinery foundations. On the positive side, it is pourable into forms
to an accurate net shape; it is relatively dense (typical density = 150 lb./ft.3), stiff (typical Young’s
Modulus = 3 x 106 psi), strong (in compression - with typically achievable compressive strength of
4,000 psi) and durable; it has a coefficient of thermal expansion close to that of steel.

6 . 2 . 2 Tensile Strength of Concrete

Concrete’s tendency to crack under tensile stress represents its primary negative characteristic.
While manageable, this tendency requires special attention in the context of the static and dynamic
loads prevailing in a reciprocating compressor foundation. Concrete strength is normally
quantified in terms of compressive strength, but tensile strength governs concrete’s tendency to
crack. The finite element analyses discussed in these guidelines used the tensile

6-1
strength = 6 * (compressive strength)0.5. In addition, dynamic loads repeating typically 18,000
times per hour make fatigue a concern - essentially reducing by an additional factor of two the
tolerable tensile stress.

The above formula for tensile strength, with 4000 psi compressive strength concrete, yields a
tensile strength of 380 psi, downgradeable under dynamic loads to 190 psi, due to fatigue
potential.

6 . 2 . 3 Evolving Characteristics

The technology of concrete is evolving, so guidelines and cautions written today may be less
relevant for blocks installed in a few years time. In particular it appears possible, with such
additives as polymer and silica fume, to achieve substantially higher tensile strengths than those
achievable with conventional concretes, but this is not widely practiced. In general, if a concrete
with predictable tensile strength of 1000 psi were routinely available, without any undesirable side
effects of this added tensile strength, the problem of cracking under tensile loads could be readily
eliminated with appropriate engineering.

6.3 Rebar

Rebar reinforces the concrete and gives it increased tolerance to tensile stresses. Previous practice
was to use less rebar than is normally placed in today’s foundations. Some practitioners now
recommend as much as #8 (1 inch diameter) rebar on 8 inch centers, which gives a rebar density of
just over 1.0 percent. A previous PCRC research project into foundation parameters shows that
the presence of rebar in the range 0.2 percent to 1.0 percent by volume significantly reduced the
extent of cracking around the anchor bolt termination point but did not eliminate it. Management of
tensile stresses in the concrete and concrete's tensile strength are clearly necessary to inhibit
cracking, but rebar has a significant role to play in the integrity of the installation by its ability to
control the extent of cracking once it occurs. Thus, a high density of rebar (as much as 1 percent
or above) appears valuable in reciprocating compressor installations.

6-2
6.4 Foundation Geometry

6 . 4 . 1 Shape

Reciprocating compressor foundations tend to have a rectangular shape configured to provide


support for the compressor frame, the crosshead guides, and in many cases, the discharge bottles
of the compressor.

6 . 4 . 2 Stress Concentration

Geometries which may cause stress concentration should be avoided.

In particular, designs which provide concrete only below the frame and crosshead guides, with
“re-entrant” spaces in between side-by-side crosshead guides should be avoided. In horizontally
opposed configurations, such a geometry has shown a tendency to split down the middle!

6 . 4 . 3 Reducing Stress Concentration in the Pan Recess

The space under the oil pan experiences alternating stress, and its geometry should seek to
minimize the stress level, and avoid sharp corners at the bottom of the oil pan “trough”. Analysis
and crack patterns suggest these corners are subject to the highest dynamic stresses.

6 . 4 . 4 Controlling Nominal Tensile Stress

One source of high tensile stresses is the gas force transmitted through the crosshead guide tie-
downs, particularly in horizontally opposed machines. Another source is the horizontal shaking
forces transmitted through the frame tie-downs. Analysis suggests that a wide flat expanse of
concrete, extending as far as possible outboard of the pan recess helps to control the magnitude of
tensile stresses induced in the pan recess area. Making the pan recess as shallow as possible also
tends to reduce stress sensitivity to imposed loads.

6 . 4 . 5 Interfaces

Avoid or minimize “vertical interfaces” between stiff materials. For example, sole plates embedded
in epoxy with sharp unrelieved corners tend to develop cracks (see Figure 6-1 for a particular
example of observed crack patterns). Use of grout “pockets” to support chocks or sole plates with

6-3
vertical interfaces to concrete have natural tendency to develop stresses in the concrete, some of
them tensile, because the epoxy has a much higher coefficient of thermal expansion than the
concrete.

6 . 4 . 6 Location of Anchor Bolts

Avoid locating the preloaded anchor bolts and their terminations near to the vertical outer surfaces
of the foundation. Any cracks at the bolt terminations will tend to grow towards the vertical
surface, create visible cracking at this location, and create a potential point for oil entry.

6.5 Loads and Concrete Stresses

6 . 5 . 1 General

The dynamic horizontal loads transmitted through the tie-downs and the vertical loads applied at the
termination of preloaded anchor bolts are of most concern in stress analysis of reciprocating
compressor foundations.

6 . 5 . 2 Termination Loads

Providing sufficient anchor bolt preload to hold dynamic loads by friction produces tensile stresses
near the bolt termination that make cracking difficult to avoid with typical concrete tensile strength.
The most promising methods to avoid cracking at the bolt termination in the block are:

• Use concrete with tensile strength in the region of 1,000 psi.


• Move the termination point from the foundation block into the mat.

The first of these two methods still remains a challenge for current technology; the second is
achievable with current technology, practice, and experience.

6 . 5 . 3 Horizontal Loads and Stresses

The dynamic horizontal loads, transmitted to the block through sound tie-downs, cause alternating
stresses at the corner of the pan recess. The tensile component can cause cracking at this location.
Geometry tends to influence the severity of these stresses. To minimize stresses:

6-4
• use rounded corners at the bottom of the recess;
• make the recess as shallow as possible;
• extend the concrete as far as possible to carry the applied horizontal load;
• install vertical post-tension bolts close to the corner of the pan recess to reduce this tensile
stress component by 100 psi or more (see Figure 6-2) based on PCRC research. In many
cases, this represents a significant reduction relative to concrete’s tensile strength of 300 to 400
psi.

6.6 Calculation and Assessment of Stresses

6 . 6 . 1 Finite Element Methods

A finite element analysis of the concrete block subject to compressor loads and to the anchor bolt
preload will provide the most reliable basis for determining the severity of stresses under loads.
Currently available finite element programs have special concrete elements. The ANSYS program
has been used in PCRC research, and its concrete element allows specification of concrete strength
which is different in tension and compression. The program treats cracking with a non-linear
criterion. If the program identifies a cracked element, it recalculates the stresses with the load
capacity of that element reduced. If it finds new elements with cracks under this new condition it
repeats the process. The program continues to grow the extent of the cracked region until it finds
no new cracked elements in the last evaluation. Even with such a model the local detail needed to
make accurate stress predictions normally limits the extent of the model - perhaps to a half
symmetry model of the block with axial extent to cover one cylinder. As PC capacity and speed
grows, such limitations will become less necessary. Appendix B summarizes a finite element
analysis performed on a large reciprocating compressor foundation.

6 . 6 . 2 Bolt Termination Stresses

With concrete’s normal tensile strength, cracking will tend to occur at bolt terminations in the
foundation block. The only available methods of inhibiting this tendency have been identified
above (higher strength concrete, or moving the termination point into the mat). Such cracks at the
termination should not threaten the structure unless they grow out to the vertical surface and/or get
penetrated by oil.

6-5
6 . 6 . 3 Estimating Pan Recess Stresses

One method of calculating stresses in the pan recess area, using a simplified strength of materials
approach, treats the loaded portion of the block above the bottom of the pan recess as a “stubby”
beam. This beam will act somewhere between a guided cantilever (no rotation at its end) and a
beam free to rotate at its end. Figure 6-3 shows the configuration. Based on this configuration the
recess area stress can be estimated as follows:

Stress = Ksc * M * y/I (6-1)

3
= Ksc * Kb * F * Hr * (d/2) * 12 / (b * d )

where:

Ksc = a stress concentration factor to account for the corner of the pan area.
Kb = a factor between 0.5 and 1.0 to account for the rotational freedom at the top; a value
of 1.0 for K b is an appropriate, conservative, choice in the absence of other
information.
M = the bending moment at level of the base of the pan recess induced by the applied
force.
F = the applied force (load).
y = the half section extent of the assumed beam (measured parallel to the load).
I = the second moment of area of the assumed beam.
Hr = the height from the bottom of the pan recess to the chock or soleplate.
b = the width of the assumed beam (perpendicular to the load); “cylinder” width at the
crosshead guide tie-downs is a conservative choice for b.
d = the section extent of the assumed beam (dimension parallel to load).
r = the “fillet radius” at the corner of the pan recess.

Peterson (Ref. 17) provides a source of stress concentration data. Some interpretation of
Peterson’s Figure 72 for bending of a stepped, flat bar gives Ksc as a function of (r/d) (see
Table 6-1). The range of 2 to 3.3. for Ksc in this table emphasizes the increased integrity which a
generous radius at the bottom of the pan provides as protection against the commonly observed
cracks initiating at this point.

6-6
If the stress so calculated exceeds half of the tensile strength of the concrete, where half represents
a fatigue strength reduction factor under dynamic load, some further attention is needed, such as:
reduction of load, change in geometry, increase in concrete strength, or more detailed analysis.
The assumptions in this analysis appear to be conservative, but are generally expected to yield
geometrical requirements which are achievable.

Example:

For a gas rod load of 108,000 lb., estimate pan stress with a pan recess height of 17 inches (Hr), a
cylinder width (b) of 24 inches, 60 inch section extent (d); assess fatigue potential with 4,000 psi
compression strength concrete, applying a factor of 2 load reduction, with a fillet radius of 1.5
inches at the bottom inside corner of the pan recess.

60
Ksc * 54,000 * 17 * * 12/(24 * 603)
2

= Ksc * 63.75

Extracting Ksc = 2.54 for r/d = .025 from Table 6-1 yields a dynamic stress of 162 psi.

An estimate of fatigue strength is:

0.5 * 6 * (compressive strength)0.5

= 190 psi

where:

0.5 = the fatigue strength reduction factor.

The stress level of 162 psi compared to the strength of 190 psi appears satisfactory for fatigue,
with 4,000 psi compressive strength concrete, recalling that the analysis is based on conservative
assumptions for load and stress.

6-7
6.7 Post-Tensioning

Post-tensioning, as most commonly practiced in reciprocating compressor foundations, is a repair


technique. The strength of cracked concrete is restored by forcing the cracked interface into
compression by long rods, acting in tension, each passing through a hole in the concrete, and
pulling together the two sections divided by the crack. The approach has often successfully
restored integrity to a weakened block.

There has been limited application of post-tensioning as a means to increase the effective tensile
strength of the concrete as part of the design. The approach is to place tie rods in two or three
orthogonal directions with tension designed to impose an average compressive stress on the
concrete of about 100 psi. This will tend to offset any tensile stresses which develop. To
accomplish this, high tensile strength rods are left in place with threads at each end; around each
rod PVC pipes are placed prior to the pour. After curing for enough time to build compressive
strength, the rods are tensioned against plates embedded in the side of the block. Rods are usually
placed on two foot centers near the top of the block. While not ensuring against cracking, this
helps raise the threshold at which cracks occur, and will certainly inhibit the growth of cracks.
Those applying this technique speak highly of it and report good experience. Figures 6-4a and 6-
4b show an example of a concrete foundation block post-tensioned in this way as part of the
design. Although not shown in these figures, this block had the normal recess for the compressor
pan. The operator reports successful implementation of this configuration of post-tensioning, and
no cracks in operation. Measured increase in compressive stress matched predictions.

PCRC research, using finite element methods, confirms it is possible to reduce tensile stress by
about 100 psi near the pan recess. Two post-tensioned rods are installed near the recess and
tensioned to approximately 100,000 lb. (as previously discussed relative to Figure 6-2).

6-8
TABLE 6-1. VALUES FOR STRESS
CONCENTRATION FACTOR, Ksc, INFERRED
FROM REFERENCE 17
r/d K sc
.01 3.3
.015 2.95
.02 2.7
.025 2.54
.03 2.4
.035 2.3
.04 2.2
.05 2.07

6-9
Force
Compressor Oil Pan
Cylinder
C
L Grout
Pockets
Sole Plate

Rail

Typical Crack
Patterns

FIGURE 6-1. FOUNDATION BLOCK ISOMETRIC VIEW - COMPRESSOR


AREA SHOWING GROUT POCKET, SOLE PLATES AND
OBSERVED AREAS OF CRACKING

250

200
2 Horizontal Bolts
Principal Tensile Stress, psi

150

2 Vertical Bolts
Near Pan
100

50

0
0 20000 40000 60000 80000 100000 120000 140000
Total Bolt Load, lb.

FIGURE 6-2. PREDICTED INFLUENCE OF HORIZONTAL AND VERTICAL


BOLTS ON TENDENCY OF CONCRETE TO CRACK IN TENSION NEAR PAN

6-10
b = axial width of
F one cylinder

Hr r
d

FIGURE 6-3. CONFIGURATION FOR ESTIMATING STRESS


IN THE PAN AREA

6-11
FIGURE 6-4a. EXAMPLE BLOCK WITH LATERAL AND LONGITUDINAL
POST-TENSIONING RODS
(ILLUSTRATIONS COURTESY OF PANENERGY CORP.)

6-12
FIGURE 6-4b. TERMINATION DETAILS FOR POST-TENSIONING RODS
(ILLUSTRATIONS COURTESY OF PANENERGY CORP.)

6-13
7.0 INSTALLATION

An experienced installer of reciprocating compressor foundations is essential. Many important


details of practice, if not addressed, can negate most of the benefits of good design. When the
design has specific features chosen for long term integrity, it is important that the installer
understands these, agrees with them, and will ensure the installation realizes them. A good quality
program is important. Ideally, a company or individual possessing engineering experience with
reciprocating compressor foundations, and understanding these guidelines, would prepare the
drawings for the installer to bid on and to use in executing the installation.

This document does not purport to provide broad experience relative to installation. Such
experience is critical, however, and should be carefully sought, by reference and past successful
examples.

A few items relevant to installation appear in the guidelines, as follows:

• The arguments for aligning the compressor level have been presented in Section 4.6.4.1.
• The arguments for controlling stress in the pan recess by ensuring a generous radius have been
presented in Section 6.4.3.
• The arguments for a rebar density as high as 1 percent have been presented in Section 6.3.
• The arguments for maximizing anchor bolt length, ideally extending anchor bolts more than
half way into the mat, have been presented in Section 5.4.
• Section 4.6.6 presents the arguments in favor of using a release agent between poured epoxy
and materials with which it must interface.

In addition, some companies now specify a “drip tray” to be installed under the pan, so that oil
drips in this area will not touch the concrete; others also add protective shields (“awnings”) over
each bolt to deflect oil from dripping on the bolt. Yet others install a gutter around the block to
localize any oil and allow it to be channeled away.

7-1
8.0 OPERATION AND MAINTENANCE

Once installed, the essentials are to keep the anchor bolts tight; to minimize the exposure of the
foundation and mounts to oil; to periodically inspect and monitor condition as discussed below.
Hopefully good design, good installation, and good operating practices will ensure the service life.
These guidelines and the experience of a good installer represent many lessons learned from
observations of the penalties for inadequate design and installation. With current knowledge and
good operating practice, there is no reason not to expect the foundation to last as long as the
compressor is needed.

8-1
9.0 INSPECTION AND CONDITION MONITORING

The operator of the compressor installation should follow good practice in inspecting and assessing
the foundation condition and taking action when required. The following are important to the
integrity of the foundation:

• Check anchor bolt tension regularly and tighten immediately if out of range.
• Check anchor bolts every few months in the first year after installation and annually thereafter.
• Check web deflections at least once per year.
• If web deflections are out or cannot be corrected by other means (e.g., ensuring consistent
bearing clearance and thickness), realign the machine using the adjustable chocks.
• Monitor vibration of the compressor at the bearing centerline. Ideally, vibrations will be less
than 10 mils.
• If vibrations are higher than ten mils they should be regularly monitored, and trended.
• If vibration are growing steadily over several surveys or take a sudden unexplained jump, the
situation and potential explanations should be careful reviewed.

9-1
10.0 SUMMARY OF GUIDELINES

1. Seek to make the compressor and foundation an integrated structure.


2. Recognize the installation as a system.
3. Recognize responsible parties.
4. Understand how each party contributes.
5. Be aware of how responsibilities overlap.
6. Ensure responsible parties communicate.
7. Apply guidelines to large slow speed reciprocating compressors.
8. Apply guidelines to achieve mechanical integrity of compressor installation.
9. Use other sources (Refs. 10 through 12) for the important topic of foundation vibration upon
soil.
10. Guidelines should ensure foundation locates and supports compressor, keeps it aligned and
is able to carry applied loads over its life.
11. Guidelines address applied forces, force management, force transmittal, mount integrity,
anchor bolts, and block integrity.
12. Get maximum cylinder gas force using Equations 2-4 and 2-5.
13. Get maximum throw inertia force using Equation 2-8.
14. Get summed forces and moments using Appendix A.
15. Use throw and cylinder maximum forces as basis for loads on block mounts, and anchor
bolts.
16. Use summed forces and moments only in managing vibration response of block upon soil.
17. Ensure no significant resonant response of the block to any summed forces or moments.
18. Use Equations 3-1, 3-2, 3-3, and 3-4 to quantify:
• Transverse force applied per cylinder to block.
• CHG transverse force per bolt.
• Frame transverse force per bolt.
19. Engineer anchor bolts and mounts on basis of largest of CHG and frame forces (F max).
20. Select an adjustable chock configuration and material (steel, epoxy, or composite material).
For its friction coefficient, calculate minimum tensile anchor bolt force to carry Fmax, using
Equation 4-2 with Wr = 0.
21. Size chock to carry bolt tension with acceptable load per unit area for the material.
22. Ensure high strength anchor bolt of diameter set by frame can provide needed minimum
tensile force.
23. Maximize anchor bolt length - ideally to halfway into mat.

10-1
24. Choose concrete with at least 4000 psi compressive strength and 380 psi tensile strength.
25. Design block transverse dimension and pan recess geometry to carry applied transverse force
with acceptable stress relative to fatigue strength of concrete (190 psi minimum).
26. Get pan recess stress using Equation 6-6.
27. Use as much as one percent rebar by volume.
28. Locate recess anchor bolt at least 12 inches from any vertical face of block.
29. Minimize use of vertical interfaces between epoxy and steel or epoxy and concrete.
30. Align compressor level.
31. With poured epoxy chocks, if used, apply a mold release agent (unless specific application
data indicates otherwise).
32. Select high strength anchor bolt material with rolled threads.
33. Use spherical self-aligning washer.
34. Consider use of Multi Jackbolt Tensioner (MJT) and/or tension monitor.
35. Relieve concrete from anchor bolt for at least 10 diameter’s depth from top of block.
36. Sleeve or tape anchor bolt over its length.
37. Do not use J or L bolt terminations.
38. Use a symmetric termination of adequate, but not excessive size.
39. When operating, check anchor bolts at least annually and tighten immediately if they lose
tension.
40. Check anchor bolts more frequently in first year of installation, or whenever they have
recently been tightened.
41. Check web deflections each year and adjust as necessary.
42. Check web deflections and frame pull-down when tightening anchor bolts.
43. Consider use of post-tensioning rods in new designs (see Section 6.7) to put concrete
initially in compression and offset tensile stresses.
44. Select organizations to engineer and install foundation block with care, based on experience
and understanding of requirements.
45. Keep bolts, nuts, and block free of leaked oil.
46. Maintain and monitor installation to ensure its integrity over time.

10-2
11.0 REFERENCES

1. Smalley, A. J., “Dynamic Forces Transmitted by a Compressor to its Foundation,”


Energy-sources Technology Conference and Exhibition, New Orleans, Louisiana, January
10-14, 1988.

2. Mandke, J. S. and Smalley, A. J., “Thermal Distortion of Reciprocating Compressor


Foundation Blocks,” ASME Paper No. 92-Pet-3, Energy-sources Technology Conference
and Exhibition, Houston, Texas, January 26-29, 1992.

3. Mandke, J. S. and Smalley, A. J., “Analyzing Thermal Distortion of Compressor


Foundation Block,” Pipe Line Industry, June 1994, pp. 21-25.

4. PCRC Technical Report (TR) 89-3 “Foundation Thermoelastic Distortion”.

5. PCRC Technical Report (TR) 92-2 “Dynamics of Compressor Skids”.

6. PCRC Technical Assessment (TA) 93-1 “Reciprocating Compressor Foundations:


Loading, Design Analysis, Monitoring and Repair”.

7. PCRC Technical Assessment (TA) 94-1 “Parameter Studies for Enhanced Integrity of
Reciprocating Compressor Foundation Blocks”.

8. Smalley, A. J., “Crankshaft Stress Reduction Through Improved Alignment Practice,


Project Final Report Incorporating Report on Field Test of a Sole Plate Mounted
Reciprocating Compressor,” A.G.A. Project PR15-174, for Pipeline Research Committee,
American Gas Association, December 1985.

9. Smalley, A. J., Pantermuehl, P. J., Lewis, R. M., and Johnson E. A., “Design of
Reciprocating Compressor Mounting Systems and Foundations for Integrity and
Reliability,” Pipeline and Compressor Research Council (PCRC) 1996 Gas Machinery
Conference, September 30-October 2, 1996, Denver, Colorado.

10. Richart Jr., F. E., Hall Jr., J. R., and Woods, R. D., Vibrations of Soils and
Foundations, Prentice-Hall Inc., 1970.

11-1
11. Arya, Suresh C., O'Neill, Michael W., and Pincus, George, Design of Structures and
Foundations for Vibrating Machines, Gulf Publishing Company, 1979.

12. Bowles, Joseph E., Foundation Analysis and Design, Fourth Edition, McGraw-Hill Book
Company.

13. “Grouting for Support of Equipment and Machines,” ACI Report 351.1R-93.

14. Baumeister, Theodore, Avallone, Eugene A., and Baumeister III, Theodore, Marks'
Standard Handbook for Mechanical Engineers, Eighth Edition, McGraw-Hill Book
Company.

15. James, R. W. and Winston, W. G., “Predicting Creep Lifetimes for Epoxy Grouts Under
Integral Gas Compressors,” ASME Transactions, April 1983, pp. 217-222.

16. “Standard Specification for Alloy-Steel and Stainless Steel Bolting Materials for High-
Temperature Service,” ASTM Designation: A 193/A 193M - 94b.

17. Peterson, R. E., “Stress Concentration Factors,” John Wiley & Sons, Inc.; ISBN 0-471-
68329-9.

18. Smalley, Anthony J., Final Report - “Compressor Diagnostics Software: Development,
Test, and Evaluation,” GRI Report No. GRI-95/0448, December 1995.

19. Smalley, A. J., Berry A. R., and Kothari, K. M., “Compressor Diagnostic Software
Package Guides Cost Control,” Pipe Line Industry, August 1993, pp. 33-40.

11-2
12.0 ACKNOWLEDGMENTS

The authors wish to acknowledge the long term Champion for the PCRC Foundation Program,
Rose Mary Lewis of Williams Natural Gas Company. She has provided encouragement and
regular feedback through the several years the project has run, and her critical, constructive review,
and editing of this document has materially added to its quality and value.

The authors also wish to acknowledge Elizabeth Johnson of Amoco Corporation, co-Champion of
the program, whose encouragement and critical review has appropriately strengthened the project
emphasis on mounting technology.

12-1
APPENDIX A

CALCULATION OF UNBALANCED SHAKING


FORCES AND MOMENTS

A-1
THE DEVELOPMENT OF THE COMPUTATIONAL METHOD

The inertial forces generated in a single-slider crank mechanism are inherently unbalanced. For
multi-cylinder engines, the resultant forces and moments acting on the foundation depend on the
arrangement of individual cylinders, their crank throw orientation (phase angle), and the actual
weights of the components of the individual cylinders. Generally, all OEM’s provide the
magnitudes of unbalanced forces and moments for their compressor units. These calculations are
based on the design weights of each component, and depending on the manufacturing tolerances,
the actual unbalanced forces may vary slightly from those computed values. When compressor
components are to be replaced, it is not always possible to obtain the replacement parts with
weights identical to the original components. In such cases, the operator should perform shaking
force calculations and add balancing weights, as needed to minimize the resulting unbalanced loads
acting on the compressor. The following provides an outline of the procedure to compute the
inertial loads in a reciprocating compressor. The details of these equations can be found in any text
book on theory of machines or machinery dynamics.

Assuming,

r = crank radius.
l = connecting rod length.
ω = angular velocity of crankshaft, rad/sec.
2πN
= , N is shaft velocity in RPM
60
Wc = total weight of rotating components.
Wr = total weight of reciprocating components.
t = time.
g = gravitational constant.

Assuming that the cylinder is located in the horizontal plane, the resulting inertial forces for a single
cylinder compressor, in the horizontal and the vertical plane are given approximately by:

rω 2 Wr r 2ω 2
Fx = [Wc + Wr ] ω
cos t + cos 2ωt (A-1)
g g l

A-2
Wc
Fy = rω 2 sin ωt (A-2)
g

Thus, the horizontal force, F x in Equation (A-1), has a primary component (first term) with the
driving shaft frequency and a secondary force component with frequency twice the driving shaft
frequency. The vertical force Fy has only the primary component.

For a multi-cylinder component unit, with ‘n’ number of cylinders, the unbalanced forces for the i-
th cylinder can be written as:

Wri ri2ω 2
[ ci ri ] i
1
Fxi = W + W r ω 2
cos (ωt + α ) + cos2 (ωt + α i ) (A-3)
l
i
g g

and,

Wci
Fyi = ri ω 2 sin (ωt + α i ) (A-4)
g

Where αi is the phase angle for the crank radius of the i-th cylinder. Noting that the above
equations have force components at one times (primary) and two times (secondary) the crankshaft
driving frequency, they can be written as:

Fxi = Fxi′ + Fxi′′ (A-5)

Fyi = Fyi′ (A-6)

[Wci + Wri ] ri ω 2 cos (ωt + α i )


1
with Fxi′ = (A-7)
g

Wri ri2ω 2
Fxi′′ = cos2 (ωt + α i ) (A-8)
g l

A-3
Wci
Fyi′ = ri ω 2 sin (ωt + α i ) (A-9)
g

In the above equations, Fxi′ and Fyi′ are, respectively, the horizontal primary and the vertical
primary components, and Fxi′′ is the horizontal secondary component of the unbalanced loads.

Resultant forces Fx and Fy transmitted to the foundation by ‘n’ cylinders are given by:

n
Fx = ∑[F′
i =1
xi + Fxi′′] (A-10)

n
Fy = ∑ F′
i =1
yi (A-11)

If all cylinders have identical geometry and component weights, Equations (A-10) and (A-11)
simplify to:

rω 2  n
r n

(Wc + Wr ) ∑ cos (ωt + α i ) + ∑ cos2 (ωt + α i )
g 
Fx = Wr (A-12)
i =1 l i =1 

n
Wc
Fy = rω 2
g
∑ sin (ωt
i =1
+ αi ) (A-13)

n n
where: Wc = ∑W
i =1
ci and Wr = ∑W
i =1
ri

Referring to Figure 1, the horizontal and the vertical moments due to the above unbalanced forces
are given by:

n
Mx = ∑F
i =1
yi Zi (A-14)

A-4
n
My = ∑F
i =1
xi Zi (A-15)

The above equations assume that all cylinder axes are located in the XZ plane. The quantities F xi
and Fyi are given by Equations (A-3) and (A-4).

If the cylinders are opposed or inclined to global x-axis with orientation angles φi, then Equations

(A-3) and (A-4) can be considered as forces in local axes with global components FxiG and FyiG
given by:

FxiG = Fxi cos φi - Fyi sin φi (A-16)

FyiG = Fxi sin φi - Fyi cos φi (A-17)

The resultant forces and moments due to ‘n’ cylinders is global coordinates can be calculated in the
same manner as described earlier. They are given by:

∑ [ F ′ cos φ ]
n
FxGP = xi i - Fyi′ sin φi (A-18)
i =1

n
FxGS = ∑ F ′′ cos φ
i =1
xi i (A-19)

∑ [ F ′ sin φ ]
n
FyGP = xi i - Fyi′ cos φi (A-20)
i =1

n
FyGS = ∑ F ′′ sin φ
i =1
xi i (A-21)

∑ [ F ′ sin φ ]
n
M xGP = xi i - Fyi′ cos φ i Zi (A-22)
i =1

n
MxGS = ∑ [ F ′′ sin φ ] Z
i =1
xi i i (A-23)

A-5
∑ [ F ′ cos φ ]
n
MyGP = xi i - Fyi′ sin φi Zi (A-24)
i =1

n
MyGS = ∑ [ F ′′ cos φ ] Z
i =1
xi i i (A-25)

where:

FxGP = horizontal primary force.


FxGS = horizontal secondary force.
FyGP = vertical primary force.

FyGS = vertical secondary force.

M GP
x = horizontal primary couple (about x-axis).
M GS
x = horizontal secondary couple (about x-axis).
M GP
y = vertical primary couple (about y-axis).

M GS
y = vertical secondary couple (about y-axis).

When counterweights on the crankshaft are used to reduce the resulting unbalanced forces or
moments, their effect can be included by using Equations (A-5) through (A-9) and using Wr = 0
and Wc to represent the actual weight of the counterweight. Then, these forces can be included into
Equation (A-18) through (A-25) to calculate the resultant unbalanced forces and moments. Note
that the rotating counterweight results only in primary forces and couples.

The terms Wri and Wci in Equations (A-3) and (A-4) include component weights as defined below:

• Reciprocating Weights (Wri):


- Piston
- Piston Rod

A-6
- Crosshead
- Crosshead Pin
- Crosshead Nut
- Cover Plates
- Connecting Rod Small End
- Tie Bolts, Nuts, Washers, Etc.

• Rotating Weights (Wci):


- Connecting Rod Big End
- Crank Pin
- Washers
- Counterweights
- Unbalanced Part Crank Web

Generally, the crankshafts are forged and machined to check complete balance. Crank webs are
symmetrical relative to the shaft axis and with accurate weight balance for each web. However, if
this not the case, and if the unbalanced equivalent weight at the crank pin location can be
computed, then it can be included in the rotating weight, Wci.

Equations (A-18) through (A-25) provide instantaneous values of the various force and moment
components for a given value of ωt. For design purposes, it is necessary to compute the peak-to-

peak values of these components. By computing the primary forces and moment for ωt = 0 and 90

degrees and computing secondary forces and moments at ωt = 0 and 45 degrees, it can be shown
that the amplitudes of the various force and moment components are given by:

FH1 = FXGP = [(F )


P 2
x0 + ( FxP90 ) ]
2 1/ 2
(A-26)

FV1 = FyGP = [(F )


P 2
y0 ( )]
+ FyP90
2 1/ 2
(A-27)

FH 2 = FxGS = [(F )
S 2
x0 + ( FxS45 ) ]
2 1/ 2
(A-28)

A-7
FV 2 = FyGS = [(F )
S 2
y0 ( )]
+ FyS45
2 1/ 2
(A-29)

MH1 = MxGP = [( M )
P 2
x0 + ( MxP90 ) ]
2 1/ 2
(A-30)

MV1 = MyGP = [( M )
P 2
y0 (
+ MyP90 )]
2 1/ 2
(A-31)

MH 2 = MxGS = [( M )
S 2
x0 + ( MxS45 ) ]
2 1/ 2
(A-32)

MV 2 = MyGS = [(M )
S 2
y0 (
+ MyS45 )]
2 1/ 2
(A-33)

where:

FxP0 , FxP90 = horizontal primary forces for ωt = 0, 90.

FyP0 , FyP90 = vertical primary forces for ωt = 0, 90.

Fxs0 , FxS45 = horizontal secondary forces for ωt = 0, 45.

FyS0 , FyS45 = vertical secondary forces for ωt = 0, 45.

MxP0 , MxP90 = primary moments about x-axis for ωt = 0, 90.

MyP0 , MyP90 = primary moments about y-axis for ωt = 0, 90.

MxS0 , MxS45 = secondary moments about x-axis for ωt = 0, 45.

MyS0 , MyS45 = secondary moments about y-axis for ωt = 0, 45.

When calculating shaking forces, the gas loads are generally assumed to cancel each other and as
such are excluded from the calculation of unbalanced loads. However, this assumption may not be
valid if the main frame has significant flexibility and the unbalanced loads are to be used for sizing
the anchor bolts that tie down the frame to the skid.

A-8
The unbalanced forces and moments generated due to inertial loading in a compressor can be
computed by hand using the procedure and the equations given so far. If such computations are
needed routinely on several different units, it will be easy to prepare a spreadsheet to input the
necessary data for each machine, and to let the program compute the necessary components of
primary and secondary forces and moments. Such spreadsheets can be prepared on a personal
computer by using any spreadsheet program that the user prefers.

ANALYSIS FOR COUNTERWEIGHTS

Counterweights are often added to the crankshaft as a means of offsetting unbalance, and as a
means of controlling the summed forces and moments. Since counterweights rotate at shaft speed
and do not reciprocate, they can only contribute to forces acting at first order of rotational speed,
ω.

The individual points of attachment for counterweights are normally the webs of each throw,
whose axial location differs a small amount from the center of the adjacent throw. A
comprehensive analysis of the contributions of counterweights to the summed forces and moments
would consider each counterweight (j) have the following attributes:

(Wwrw)j = unbalance * radius of unbalance vector (inch-lb.).


Ψwj = angle of unbalance vector from reference throw.
Zwj = axial location of counterweight line of action, usually referred to
unit mid-point.
Nw = total number of counterweights to be considered.

Adapting Equations A-7 and A-9, we obtain:

Fxwj = (Wwrw)j ω2 cos (ωt + Ψwj)

Fywj = (Wwrw)j ω2 sin (ωt + Ψwj)

Nw
Fxw = ∑
j =1
Fxwj

A-9
Nw
Fyw = ∑j =1
Fywj

Nw
Mwx = ∑j =1
Fywj Zj

Nw
Myx = ∑
j =1
Fxwj Zj

where:

F xwj, Fywj = individual counterweight unbalance forces in X and Y directions.

F xw, Fyw = summed counterweight forces in X, Y directions.

Mxw, Myw = summed counterweight moments about point Z = 0.

A-10
Cylinder
No. i

Zi

Xi
X

Crankshaft
Z C
L

FIGURE A-1. CYLINDER GEOMETRY

A-11
APPENDIX B

SUMMARY OF FINITE ELEMENT ANALYSIS OF A


RECIPROCATING COMPRESSOR AND
ITS FOUNDATION

B-1
MODELING OF COMPRESSOR FOUNDATION INTERACTION

The frame and concrete block under load act as a flexible interacting structure. To properly account
for the influence of frame and concrete flexibility on transmitted loads requires a model that
convincingly addresses the main contributors to flexibility of the compressor frame, the deflection
characteristics of the concrete block, and the interaction of the two through a series of mounting
mechanisms. Figure B-1 shows such a model. It uses elements within the ANSYS finite element
package to model a complete compressor frame, the crosshead guide (doghouse), the tie-downs
and mounts on frame and crosshead guide, and the foundation block itself. The compressor frame
elements consist predominantly of plate, beam, and strut elements. The foundation block model
generally consists of 3D solid elements. This global model has the purpose of determining
deflections and transmitted forces rather than analyzing stresses. A separate more detailed model
of the foundation block and mounting mechanism provides block stresses. However, the global
model does include all significant structural details of the frame and all bolted connections between
frame and block. It also includes a simplified model of the crankshaft, spring elements to represent
the main bearings fluid film flexibility, and points of load application on the crankshaft for the
individual cylinders. Loads on the crankshaft, vertical and horizontal, are transmitted through the
shaft to the main bearings and to the tie-downs. Vertical forces from the connecting rod acting on
the crossheads produce a vertical force on the dog house which is transmitted in turn to the
crosshead guide tie-downs. The horizontal gas load acting inwards or outwards on the cylinder
causes the cylinder, distance piece, crosshead guide, and compressor frame to deform, and thereby
to load the crosshead guide tie-downs.

The compressor shown in Figure B-1 has three stages of compression. Each stage comprises two
cylinders (cylinders 4 and 6 for the first stage, 3 and 5 for the second stage, 1 and 2 for the third
stage). The software package “Compressor Diagnostics Software” (CDS) described by Smalley,
et al (Refs. 18 and 19) generated values for the gas forces. The CDS analysis accounts for piston
kinematics, real gas properties, valve dynamics and flow resistance; it matches flow from stage-to-
stage in the process of determining appropriate interstage pressures. Figure B-2 shows maximum
gas loads by stage, predicted by this software which match quite closely those provided by the
OEM. Figure B-3 shows the variation of gas and inertia forces applied to the bearings of the most
heavily loaded cylinder.

B-2
INFLUENCE OF FRAME FLEXIBILITY ON TRANSMITTED FORCES

Figure B-4 shows the variation with crank angle of horizontal crosshead guide forces calculated in
three different ways: 1) gas load variation for the most heavily loaded cylinder; 2) the transmitted
force based on a rigid frame assumption; and 3) the force transmitted to the crosshead guide bolts
using the model described in the previous section of the paper. Clearly, the modeling assumption
has a major influence on predicted maximum forces. The gas loads acting on the cylinder and the
crankshaft exceed 107,000 lb. The rigid frame assumption produces maximum forces acting on
the crosshead guide tie-downs of about 3,700 lb. The finite element modeling approach, which
accounts for relative flexibility of frame and block, produces forces of about 39,000 lb. Thus, the
stiffness of the compressor structure reduces the transmitted load to about 36 percent of the gas
load. However, a rigid frame would essentially carry all of the gas load, reducing the calculated
transmitted force in this case to 3.5 percent of the gas load - an unrealistic and grossly non-
conservative assumption.

Figure B-5 shows the forces transmitted to the frame bolts due to inertia loads alone as a function
of modeling assumption (full inertia loads, rigid frame assumption, flexible frame assumption).
Again, the modeling assumption makes a very substantial difference to the transmitted inertia load.
The rigid frame data used a very high Young's Modulus in the finite element model.

Figure B-6 illustrates a slightly different issue from the effects of frame flexibility on total
transmitted force: the maximum force transmitted at any individual tie-down. The force obtained
by dividing the total gas load (107,000 #) by the number of bolts (4), gives about 27,000 lb. The
flexible frame analysis which accounts for uneven load share gives a maximum tie-down force of
10,740 lb. (40%). The rigid frame assumption gives 1,090 lb., again a gross underestimtae.

Table B-1 summarizes the predicted behavior of the subject compressor and foundation block, and
the influence of frame flexibility in terms of forces and percentages using individual cylinder, gas,
and inertia forces as a reference. Figure B-7 makes the same comparison in bar graph format.

B-3
TABLE B-1. SUMMARY INFLUENCE OF MODELING ASSUMPTION
COMPARED TO CONSERVATIVE ASSUMPTIONS
FOR PREDICTION FORCES
Conservative Force Flexible Rigid Frame Ratio of Ratio of
Assumption Resulting Frame Results (lb.) Flexible Rigid Frame
from Results (lb.) Frame Results to
Conservative Results to Flexible
Assumption Conservative Frame
(lb.) Assumption
Force Transmitted to 107,400 38,734 3,675 36% 9.5%
Block Through
Crosshead Guide Tie-
downs = Maximum Gas
Load/Cylinder
Individual Bolt Force for 26,847 10,740 1,090 40% 10.1%
Crosshead Guide Tie-
down = Total Gas Load /
Number of Bolts
Force Transmitted to 38,000 7,747 578 20% 7.5%
Frame Bolts Due to
Inertia Loads = Total
Inertia Load

Running a detailed flexible frame analysis has the benefit of reducing conservatism. In the
particular example, this reduction in conservatism is significant, but not nearly as dramatic as the
underestimate using the non-conservative rigid frame assumption. If anchor bolt strength and
mount coefficient of friction can generate sufficient holding power, the conservative assumption
has much to commend it. Simple, with a built-in safety factor, it avoids the need to prepare,
execute, and interpret a detailed finite element model. The finite element approach, by relaxing
conservatism, adds value when the selected tie-down system cannot achieve the required hold
down capacity, based on this conservative assumption. Once a finite element model has been
generated for a particular compressor frame configuration, subsequent analyses for different
compressor loads on that particular frame configuration are more cost effective.

B-4
The preceding graphics and table emphasize the inappropriateness of designing a foundation and
tie-downs from summed shaking forces and moments. The most essential point is to base tie-
down size and strength analysis and block integrity analysis on individual cylinder and throw
forces, not on the summed shaking forces and moments (rigid frame assumption).

Response of the Concrete Block to Applied Loads

The loads transmitted through the compressor mounts and anchor bolts cause distortion and stress
of the foundation.

The finite element model in Figure B-8 represents a section of concrete foundation block - with
sufficient detail for calculating stresses in response to applied loads. Two primary concerns exist:
the effect of anchor bolt loads near their point of termination; and the horizontal loads and their
effect at high stress concentration points such as the corner of the recess under the compressor pan.
In Figure B-9, the light colored section near the termination for each anchor bolt indicates the
predicted tendency of the block to develop negative stresses here. In a large block such as this
with substantial amounts of reinforced concrete between the bolt terminations and the outer
surfaces of the block, such cracks should probably not cause major concern. The anchor bolt
maintains its tension as a result of a compression cone above the termination point. However, if
the anchor bolt comes close to the outer surface of the block at any point, such cracks have the
potential to penetrate to the wall of the block; examples abound.

Parametric studies have shown various means of reducing the tendency to develop cracks in the
block at the bolt termination by:

1) Increasing the tensile strength of the concrete.


2) Lengthening the anchor bolt so that it passes all the way through the block and terminates
half way into the mat.
3) In repair situations, supplementing the original low strength anchor bolt using a sleeve
section with projections at 45 degrees, embedded in new high strength reinforced concrete
or epoxied into the old concrete (see Figure 5-2 in the main body of the guidelines).

Parametric studies have shown that a tensile strength of 1,000 psi, as opposed to approximately
300 to 400 psi typical of concretes with compressive strengths of 3,000 to 4,000 psi, will eliminate
this tendency to crack. Moving the termination point into the mat does not eliminate the tendency

B-5
to crack, but cracks occurring in the mat are less likely to be aggravated by the presence of
excessive oil and cannot cause side wall cracks in the block.

High stress concentrations at the corner of the recess under the compressor pan cause concern from
a structural point of view. If the horizontal loads transferred to the block by inertia and gas forces
produce excessive shear loads in sections immediately below the load application points, these
shear loads may cause crack inducing tensile stresses in the pan area. Figure B-10 is a stress
contour plot showing high tensile stresses in this area. In this particular case, the substantial top
surface area and structural width of the block, as evidenced by Figure B-8, yields a pan area tensile
stress of only 90 psi - a stress unlikely to cause cracking in this instance. In general, these
horizontal loads vary at about 5 Hz for a typical slow speed reciprocating compressor, and can
accumulate 18,000 cycles an hour and a million cycles in two or three days. It is appropriate to
apply a fatigue strength derating factor of one-half to the tensile strength of the concrete in this
region. Consequently, the concrete becomes more vulnerable to any high time varying tensile
stresses such as might occur in a more highly loaded or less substantial foundation block.

In summary, this particular foundation can manage the vertical load transmitted by anchor bolt
tension and by horizontal forces transmitted through the mounts.

B-6
FIGURE B-1. MODEL OF BLOCK AND FRAME

120
108.4 107.3
100 96.3
Load in Lbs. (1,000's)

80

60

40

20

0
Stage 1 Stage 2 Stage 3

FIGURE B-2. MAXIMUM GAS FORCE ROD LOAD FOR


3 STAGE COMPRESSOR

B-7
150000

100000
FORCES, LBS.

50000

0
0 50 100 150 200 250 300 350
-50000

-100000

-150000

CRANK ANGLE (CYL #1 REFERENCE)


CYL HEAD CH VERT MAIN VERT MAIN HORIZ

FIGURE B-3. VARIATION OF GAS FORCES FOR CYLINDER #5

GAS LOAD

100000 FLEXIBLE FRAME


80000 (FE) & BLOCK

60000
40000
20000
0
-20000 0 30 60 90 120 150 180 210 240 270 300 330
-40000 RIGID FRAME

-60000
-80000
-100000
-120000

GAS LOADS FLEXIBLE FRAME (FE) & RIGID FRAME


BLOCK

FIGURE B-4. COMPARISON OF GAS LOAD TO CROSSHEAD GUIDE LOADS


BASED ON FLEXIBLE FRAME (FE) ANALYSIS &
RIGID FRAME ASSUMPTION

B-8
MAIN BEARING
INERTIA LOAD

40000

FLEXIBLE FRAME
20000 (FE) & BLOCK
LOAD IN LBS.

0
0 30 60 90 120 150 180 210 240 270 300 330
RIGID FRAME
-20000

-40000

CRANKSHAFT INERTIA FLEXIBLE FRAME & RIGID FRAME


LOAD BLOCK (FE MODEL)

FIGURE B-5. TRANSMISSIBILITY EFFECTS: COMPARISON OF INERTIA


LOAD ON CRANKSHAFT TO FRAME BOLT LOADS PREDICTED FOR
FLEXIBLE BLOCK/FRAME & RIGID FRAME

26,847
Max Cylinder Load
Divided by Number of Bolts
18000 Per Cylinder on Frame &
Cross-head Guide
16000
14000 10,740
12000
Max Individual
10000 Anchor Bolt Load

8000
Max Individual
6000 Anchor Bolt Load

4000 1090

2000
0
RIGID FLEX SOFT

FIGURE B-6. ANCHOR BOLT LOADS FOR RIGID, FLEXIBLE, &


SOFT FRAMES

B-9
120,000

100,000

80,000 Conservative
Force, lb.

60,000 Flexible Frame


Results (lb.)
Rigid Frame Results
40,000 (lb.)

20,000

0
Block CHG Frame
Force Bolt Bolt
Force Force

FIGURE B-7. COMPARISON OF RIGID FRAME, FLEXIBLE FRAME, AND


REALISTIC FORCES FOR RECIPROCATING COMPRESSOR FOUNDATIONS

FIGURE B-8. MODEL OF FOUNDATION BLOCK FOR


LARGE PROCESS COMPRESSOR

B-10
FIGURE B-9. PREDICTED EXTENT OF CRACKING AT
ANCHOR BOLT TERMINATION

B-11
FIGURE B-10. TENSILE STRESS CONTOURS IN PAN RECESS

B-12
APPENDIX C

FIELD DATA REPORT ON


PULL-OUT TESTS

C-1
MEMO: 10/23/95

FROM: Ralph Harris/Joe Pantermuehl

TO: Tony Smalley / Pete Harrell

COPY TO: Elizabeth Johnson


Amoco Corporation
3700 Bay Area Blvd.
Houston, Texas 77058

FAX: (713) 212-1616

SUBJECT: Summary of Anchor Bolt Tests


Project 04-8000-604

In a cooperative effort between Southwest Research Institute, Amoco, and El Paso Natural Gas,
pullout tests were performed on October 17 and 18, on compressor anchor bolts at Kutz station
near Farmington, New Mexico. The tests were performed on old foundations which had
previously supported Cooper GMVH8 F reciprocating compressors, and were in the process of
demolition. The anchor bolts were one and one-quarter inch in diameter, and consisted of two
types: (1) low strength J-bolts, and (2) high strength bolts with plate termination’s embedded in
the concrete. El Paso supplied the hydraulic ram and personnel to operate the ram, and SwRI
supplied the test instrumentation and computer data acquisition system. Elizabeth Johnson of
Amoco aided in the data collection. The data was collected using a load cell and LVDT
displacement transducer, as shown in Figure C-1.

The anchor bolts were labeled as Block 1-Bolt 1, Block 1-Bolt 2, etc. For select bolts, load versus
deflection plots are shown in Figures C-2 through C-7. In all of the cases for the low strength
bolts, failure was either pullout of the bolts (J-bolts) and/or yield of the bolt material, rather than
crushing or cracking of the concrete. During testing of the high strength bolts, the hydraulic ram
itself experienced a failure at 135,000 pounds, and no further testing was performed thereafter.

C-2
C-3
C-4
C-5
C-6
C-7
C-8
C-9
APPENDIX D

GLOSSARY

D-1
GLOSSARY

Reciprocating Compressor A positive displacement compressor which uses reciprocating


pistons to raise the pressure of the gas.

Chock A pad - typically rectangular in shape and 1 to 2 inches thick


which supports the compressor frame at each point where an
anchor bolt is applied.

Cross-head A device, driven by a connecting rod, which slides on a linear


motion bearing, and transmits horizontal motion to the piston
without side loads (which are absorbed by the slider bearing).

Cross-head Guide A box-like structure, typically of cast iron, attached to the


compressor frame which carries the cross-head bearing, and
thereby guides the cross-head.

Foundation A heavy structure, typically of concrete, which supports the


compressor.

Foundation Block The block of concrete which functions as the foundation.

Separable A reciprocating compressor whose drive comes from a separate


machine - typically engine or motor.

Integral A reciprocating compressor whose drive comes from an engine


which shares a supportive structure with the compressor
cylinders.

Anchor Bolt Bolt which passes through one of many predrilled holes in the
base of the compressor frame and ties the compressor to its
foundation.

D-2
Grout Pourable material used to assist compressor mounting, including
providing a surface which to mount the unit.

Soleplate Steel plate embedded in the foundation, using grout, as a


support for the chocks.

Expansion Joint A “break” in a surface layer which allows free expansion of that
layer, and acts to limit the horizontal extent over which the layer
is in intimate contact with the foundation block.

Valve A check valve, designed to keep the cylinder closed and leak free
during compression and re-expansion motion of the piston, and
to open when necessary to allow suction gas to enter or
discharge gas to flow out.

Piston Device driven through the piston rod by the cross-head motion,
to impose motion which varies trapped volume within the
cylinder.

Piston Rod Rod which connects the cross-head to the piston.

Mounting System Combination of chocks, anchor bolts, grout, and in some cases,
plates whose function is to support the compressor.

Crank End Trapped volume within the cylinder which is on the crankshaft
(inboard) side of the piston.

Head End Trapped volume within the cylinder which is on the cylinder
head (outboard) side of the piston.

Post-Tensioning The process of applying tension to rods which pass through the
concrete in order to put the concrete in the vicinity of the rod in
compression.

D-3

Вам также может понравиться