Download as pdf or txt
Download as pdf or txt
You are on page 1of 76

SURFACE J811

REV.
AUG81
VEHICLE
400 Commonwealth Drive, Warrendale, PA 15096-0001
INFORMATION Issued 1962-06
REPORT Revised 1981-08

Superseding J811 JUN62


Submitted for recognition as an American National Standard

SURFACE ROLLING AND OTHER METHODS FOR


MECHANICAL PRESTRESSING OF METALS

Foreword—This Document has also changed to comply with the new SAE Technical Standards Board format.

1. Scope

2. References

2.1 Applicable Publications—The following publications form a part of the specification to the extent specified
herein. Unless otherwise indicated the lastest revision of SAE publications shall apply.

1. Butz, G. A., and Lyst, J. O., "Improvements in Fatigue Resistance of Aluminum Alloys by Mechanical
Surface Prestressing." Paper presented at 1961 Western Metals Congress (ASTM).
2. Gadd, C. W., Anderson, J. O., and Martin, D., "Some Factors Affecting the Fatigue Strength of Steel
Members," SAE Transactions, Vol. 63, 1955.
3. Kudryavtsev, I. V., "The Influence of Internal Stresses on the Fatigue Endurance of Steel,"
Proceedings of the International Conference on Fatigue of Metals, Institution of Mechanical
Engineers, London; ASME. New York, 1956.
4. Atkin, R. L., and Mezoff, J. G., "Development and Testing of Magnesium Alloy Wheels." Paper
presented at Third Sagamore Ordnance Materials Research Conference, December 1956, Syracuse
University Research Institute.
5. Lessels, J. M., Strength and Resistance of Metals, New York: John Wiley and Sons, Inc., 1954.
6. Cohen, B., "Effect of Shot Peening Prior to Chromium Plating on the Fatigue Strength of High Strength
Steel," WADC Technical Note 57–178, ASTIA Document #AD 130821, 1975.
7. Sigwart, H., "Influence of Residual Stresses on the Fatigue Limit," Proceedings of the International
Conference on Fatigue of Metals, Institution of Mechanical Engineers, London; ASME. New York,
1956.
8. Dugdale, D. S., "Effect of Residual Stress on Fatigue Strength," The Welding Journal, January 1959.
9. Grover, H. J., Gordon, S. A., and Jackson, L. P., Fatigue of Metals and Structures, Prepared for Bureau
of Aeronautics, Department of the Navy, Nav Aer OO-25-534, 1954.
10. Brodrick, R. F., "Protective Shot Peening of Propellers—Residual Peening Stresses," WADC Technical
Report 55–56, Part I, June 1955.
11. Horger, O. J., "Cold Working," Section 6.9, ASME Handbook, Metals Engineering—Design, McGraw-
Hill, 1953.
12. Hertz, H., Journal of Mathematics (Crelles' Journal), Vol. 92, 1881.
13. Hertz, H., Gesammelte Werke, Vol. 1, p. 155, Leipzig, 1895.

SAE Technical Standards Board Rules provide that: “This report is published by SAE to advance the state of technical and engineering sciences. The use of this report is entirely
voluntary, and its applicability and suitability for any particular use, including any patent infringement arising therefrom, is the sole responsibility of the user.”

SAE reviews each technical report at least every five years at which time it may be reaffirmed, revised, or cancelled. SAE invites your written comments and suggestions.

QUESTIONS REGARDING THIS DOCUMENT: (724) 772-8512 FAX: (724) 776-0243


TO PLACE A DOCUMENT ORDER; (724) 776-4970 FAX: (724) 776-0790
SAE WEB ADDRESS http://www.sae.org

Copyright 1981 Society of Automotive Engineers, Inc.


All rights reserved. Printed in U.S.A.
SAE J811 Revised AUG81

14. Belajef, N. M., "On the Problem of Contact Stresses," Bulletin, Institute of Engineers of Ways and
Communication, St. Petersburg, 1917. Memoirs on Theory of Structures, St. Petersburg, 1924.
15. Thomas, H. R., and Hoersch, V. A., "Stresses Due to the Pressure of One Elastic Solid Upon Another,"
University of Illinois Experimental Station Bulletin No. 212, Vol. 27, No. 46, July 15, 1930.
16. Lundberg, G., and Odqvist, F. K. G., Proc. Ingeniors Vetenskapa Akad., No. 116, Stockholm, 1932.
17. Horger, O. J., "Stressing Axles and Other Railroad Equipment by Cold Rolling," Surface Stressing of
Metals, American Society for Metals, pp. 85–142, Cleveland, Ohio, 1947.
18. Way, S., Discussion of paper by R. E. Peterson and A. M. Wahl in Journal of Applied Mechanics, Vol. 2,
No. 2, June, 1935, pp. A-69-71.
19. Horger, O. J., "Effect of Surface Rolling on the Fatigue Strength of Steel," Journal of Applied
Mechanics, Transactions, A.S.M.E., Vol. 57, December, 1935, pp. A-128-136.
20. Love, R. J., "Cold Rolled Fillets," Engineering, August 8, 1952.
21. Ford Motor Company; Manufacturing Research Office.
22. General Motors Corporation (Patent #2,357,515).
23. Industrial Metal Products Corporation.
24. International Harvester Company (Patent #2,841,861).
25. Madison Industries, Inc.
26. "Shot Peening and Other Surface Working Processes," Supplement to Metals Handbook, Metal
Progress, (ASM), July 15, 1954, pp. 104–108.
27. Almen, J. O., Mattson, R. O., and Fonda, H. E., Report on "Surface Rolling Treatment."
28. The Foote-Burt Company, Schraner Division.
29. Timken Roller Bearing Company, "Cold Rolling of Axle Fillets."
30. Almen, J. O., "Fatigue Durability of Prestressed Screw Threads," Product Engineering, April, 1951.
31. Cogsdill Tool Products, Inc.
32. Stewart, W. C., and Ellinghausen, H. C., "Examination and Test of Failed Port Tail Shaft, USS Norfolk
(DL-1)," U.S. Naval Engr. Exp. Station, R&D Report, 040007AZ(2) NSM-000-003, December 5, 1955.
33. Dugdale, D. S., "Stress-Strain Cycles of Large Amplitude," J. Mech. Phys. Solids, 1959, Vol. 7, pp.
135–142. Pergaman Press, London, England.
34. Fuchs, H. O., "Shot-Peening Effects and Specifications," ASTM Spec. Tech. Pub. No. 196, pp. 22–32.
Philadelphia, 1958.
35. Phillips, A., "Improvement of Fatigue Life of Aircraft Components by Coining," published by ASME,
1961. Paper 61-AV-35.
36. Almen, J. O., "Fatigue Loss and Gain by Electroplating," Product Engineering, Vol. 22, No. 6, June,
1951.
37. Grossman, Nicholas, "Effect of Shot Peening on the Brittle Transition Temperature," Metal Progress, p.
352. September, 1950.
38. Water, K. T., "Production Methods for Cold Working Joints Subjected to Fretting for Improvement of
Fatigue Strength," Preprint Paper No. 75, ASTM, San Francisco Meeting, 1959.
39. SAE, Shot Peening Manual, SP-84, New York, 1952.
40. Timoshenko, S., and Goodier, J. N., Theory of Elasticity, 2nd Ed. New York, McGraw-Hill, 1951.
41. Timoshenko, S., Strength of Materials, 2nd Ed., Part II: Advanced Theory and Problems. New York:
McGraw-Hill 1941.
42. Mattson, R. L., and Roberts, J. G., "Effect of Residual Stresses Induced by Strain Peening Upon
Fatigue Strength." Internal Stresses and Fatigue in Metals. Amsterdam, Elsevier, 1958.
43. Fuchs, H. O., and Mattson, R. L., "Measurement of Residual Stresses in Torsion Bar Springs,"
Proceedings, SESA, Vol. 4, No. 1, 1946.
44. Almen, J. O., "Fatigue Failures are Tensile Failures," Product Engineering. McGraw-Hill Publishing
Co., March, 1951.
45. Machine Design, Penton Publishing Co., October 20, 1958.
46. Harper, W. A., "Explosive Hardening of Steel Proves Out in Field Service," Iron Age, Vol. 185, No. 4,
Feb. 4, 1960, p. 85.
47. Dermott, R. G., "Progress in Explosive Forming", Metal Progress. November 1959.
48. Courtesy of L. G. Johnson. Research Laboratories, General Motors Corp. Unpublished.

-2-
SAE J811 Revised AUG81

49. Thum, A., and Bruder, E., "Shape and Fatigue Strength of Rod Eyes and Similar Constructional Parts,"
Deutsche Kraftfahrtforschung, No. 20, 1939, pp. 1–10.
50. Design News, Dec. 8, 1958.
51. Wise, S., "Work-Hardening Bolt Holes in Rail Ends," The Railway Gazette, April 29, 1960, pp. 511–
512.
52. "One Ball, No Errors," Machine Design. Penton Publishing Co., February, 1958.
53. Almen, J. O., "Brittle Structural Failures with Emphasis on Welded Ships," Product Engineering. April
1953.
54. Burnheim, H. "Surface Pressing of Notched Test Specimens Consisting of Cr-Mn-V Steel VCV 100,"
Luftfahrtforschung, Vol. 20, No. 1, January 20, 1943.
55. Coombs, A. G. H., Sherratt, F., and Pope, J. A., "An Analysis of the Effects of Shot-Peening Upon the
Fatigue Strength of hardened and Tempered Spring Steel." Proceedings of the International
Conference on Fatigue of Metals, Institution of Mechanical Engineers, London; American Institute of
Mechanical Engineers, New York, 1956. London, William Clowes and Sons, 1956.
56. Courtesy H. R. Neifert, Timken Roller Bearing Co.
57. Courtesy G. F. Butz, Aluminum Co. of America.
58. Courtesy C. W. Cable, Boeing Aircraft Co.

2.2 Related Publications—The following publications are provided for information purposes only and are not a
required part of this document.

Rosenthal, D., Sines, G., and Zizicas, G., "The Effect of Residual Compression on Fatigue," Welding J.,
Research Suppl., Vol. 28, 1949.
Rosenthal, D., and Sines, G., "Effect of Residual Stress on the Fatigue Strength of Notched Specimens,"
Proc. ASTM, Vol. 51, 1951.
Norton, J. T., Rosenthal, D., and Maloof, S. B., "X-Ray Diffraction Study of the Effect of Residual
Compression on Fatigue of Notched Specimens." Welding J., Research Suppl. 1946.
Surface Stressing of Metals, American Society for Metals, Cleveland, 1947; H. F. Moore, "The Problem
Defined;" W. M. Murray, "Measurement of Surface Stresses;" J. O. Almen, "Fatigue of Metals as
Influenced by Design and Internal Stresses;" O. J. Horger, "Stressing Axles and Other Railroad
Equipment by Cold Rolling;" P. R. Kosting, "Progressive Stress-Damage."
Almen, J. O., "Fatigue Weakness of Surfaces," Product Engineering. McGraw-Hill Publishing Co.,
November 1950.
Almen, J. O., "Torsional Fatigue Failures," Part I, Product Engineering, McGraw-Hill Publishing Co.
September 1951.
Almen, J. O., "Torsional Fatigue Failures," Part II, Product Engineering, McGraw-Hill Publishing Co. March
1952.
Almen, J. O., "Residual Compressive Stress Strengthens Brittle Materials," Product Engineering. McGraw-
Hill Publishing Co., July 1953.
Mattson, R. L., and Almen, J. O., "Effect of Shot Blasting on the Mechanical Properties of Steel," (NA-115),
Final Report OSRD, 3274, 4825, 6647. Washington, 1945.
Green, W. B., "How Processing Affects Bolt Fatigue Strength," Machine Design. Penton Publishing Co.,
December 1947.
Weibull, W., "The Effect of Decarburization and Other Factors on the Fatigue Strength of Roll-Threaded
Aircraft Bolts," SAAB TN 4, Svenska Aeroplan Aktiebolaget. Linkoping, Sweden, July 1952.
Buckwalter, T. V., and Horger, O. J., "Investigation of Fatigue Strength of Axles, Press-Fits, Surface Rolling,
and Effect of Size," Transactions of the American Society for Metals, Vol. 25, March 1937, p. 229.
Horger, O. J., and Maulbetsch, J. L., "Increasing the Fatigue Strength of Press-Fitted Axle Assemblies by
Surface Rolling," Journal of Applied Mechanics, September 1936, pp. A-91 to A-98.
Horger, O. J., Buckwalter, T. V., and Neifert, H. R., "Fatigue Strength of 5-1/4 in. Diameter Shafts as
Related to Design of Large Parts," Journal of Applied Mechanics, September 1945, pp. A-149 to A-
155.
Horger, O. J., and Cantley, W. I., "Design of Crankpins for Locomotives," Transactions ASME, Vol. 68,
1946, pp. A-17 to A-33.

-3-
SAE J811 Revised AUG81

Horger, O. J., "Residual Stress." Handbook of Experimental Stress Analysis, John Wiley & Sons, Inc., New
York, 1950.
Horger, O. J., "Influence of Fretting Corrosion on the Fatigue Strength of Fitted Members," Symposium and
Fretting Corrosion, Special Technical Publication No. 144, ATM, 1953, pp. 40–51.
Horger, O. J., and Lipson, C., "Automotive Rear Axles and Means of Improving Their Fatigue Resistance,"
ASTM Spec. Tech. Publ. 72, "Symposium on Testing of Parts and Assemblies," Philadelphia, 1947.
Horger O. J., "Stresses Imposed by Processing," SAE Quart. Trans. Vol. 5, No. 3, pp. 393–403, July 1951.
Horger, O. J., and Neifert, H. R., "Effect of Surface Conditions on Fatigue Properties," Surface Treatment of
Metals, ASTM, 1941.
Horger, O. J., and Neifert, H. R., "Fretting Corrosion of Large Shafts as Influenced by Surface Treatments,"
Symposium on Large Fatigue Testing Machines and Their Results, Special Technical Publication No.
216, ASTM, 1957, pp. 81–95.
Horger, O. J., and Neifert, H. R., "Correlation of Residual Stresses with Fatigue Strength of Machine
Elements and Related Phenomena," Residual Stresses in Metals and Metal Construction. Reinhold
Publishing Corp., New York, 1954, pp. 219–253.
Neifert, H. R., and Robinson, J. H., "Further Results from the Society's Investigation of Tailshaft Failures,"
Transactions, Society of Naval Architects and Marine Engineers, Vol. 63, 1955, pp. 495–550.
Fuchs, H. O., "Trapped Stresses," Machine Design. Penton Publishing Co., July 1948.
Fuchs, H. O., "Techniques of Surface Stressing to Avoid Fatigue," Metal Fatigue, edited by G. Sines and J.
L. Waisman, University of California Engineering Extension Series. McGraw-Hill Book Co., Inc., 1959.
Dolan, T. J., "Basic Concepts of Fatigue Damage in Metals," op. cit.
Peterson, R. E., "Fatigue Cracks and Fracture Surfaces—Mechanics of Development and Visual
Appearance," op. cit.
Shook, L. L., Jr., and Long, C. L., "Surface Cold Rolling of Marine Propeller Shafting." paper presented
December 6, 1957, to the Hampton Roads Section, The Society of Naval Architects and Marine
Engineers.
Egger, Walter, and Diamond, Gerald X., "Fillet Rolling," Machine Design. Penton Publishing Co., January
5, 1961.
Garwood, M. F., ZurBurg, H. H., and Erickson, M. A., "Correlation of Laboratory Tests and Service
Performance," Interpretation of Tests and Correlation with Service, ASM, 1950.
Love, R. J., and Waistall, D. N., "The Improvement in the Bending Fatigue Strength of Production
Crankshafts by Cold Rolling," Report No. 1954/2, published by the Motor Industry Research
Association, Lindley, Warwickshire, England. 1954.
Love, R. J., "Fatigue in Automobiles," Proceedings of the International Conference on Fatigue of Metals,
Institution of Mechanical Engineers, London; ASME, New York; 1956.
Siebel, E., and Gaier, M., "The Influence of Surface Roughness on the Fatigue Strength of Steels and Non-
Ferrous Alloys," translated in The Engineers' Digest. March 1957.
Bibliography on Residual Stress (SP-125), and Supplement I (SP-167), SAE.
Evaluation of Methods for Measurement of Residual Stress, (HS 147), SAE.
Horger, O. J., "Cold Working," ASME Handbook Metals Engineering-Design, 2nd edition, O. J. Horger, ed.
McGraw-Hill, 1964, pp. 264–267.
Metals Handbook, Vol. 3, 8th edition, "Roller Burnishing," pp. 105–107, "Thread Rolling," pp. 130–145.

3. Applications of Mechanical Prestressing (by George A. Butz)

3.1 Introduction—The word "prestressing" implies that a stress is applied prior to service. For the purposes of
this discussion this is true, but insufficient, definition. It must be extended to say, by virtue of a localized
pressure on the surface of a part, that the surface of the part in the vicinity is stressed in tension beyond its
elastic limit. When the pressure is removed, the surface elements tend to retain part of the total deformation
experienced under pressure. Since this is resisted by subsurface layers which did not exceed the elastic limit,
the surface and adjacent layers are left in a state of compressive residual stress.

-4-
SAE J811 Revised AUG81

The two most widely used methods of mechanical prestressing probably are surface rolling and shot peening.
Since the process of shot peening has been rather widely discussed in previously published literature, the
greater part of this manual is concerned with surface rolling and its theory, load specification, tooling, control,
and effects. Methods briefly considered include hammer peening, cold pressing, and treatment of small holes
with balls or tapered pins. The general aims of this manual are:

1. To give the reader a general understanding as to what mechanical prestressing is and whether it may
be expected to help him with his product.
2. To help him choose a process.
3. To help him get started in tool design and preparation of test samples. At this stage of development of
the art, the optimum prestressing conditions and degree of performance improvement should be
established by objective and destructive tests, unless one has ample previous experience on similar
materials and products.

Mechanical prestressing methods affect the surface layers of a part in at least three ways, the relative amounts
being affected by the process and the material:

1. Compressive residual stresses.


2. Cold work or strain hardening.
3. Surface geometry or finish.

It is usually quite difficult to assess the individual contributions of these effects on the improvement in
performance attained. The consensus, however, is that the compressive residual stress is the most potent of
the three. These effects are discussed in later sections of this report.

While mechanical prestressing methods are now used in many various industries, most of their development
and application has occurred in the transportation industry. This might be attributable to the intense
competition which fosters development work, continually driving toward the attainment of maximum strength in
minimum space and weight with low-cost alloys and processing. In meeting these goals, prestressing methods
have made some of their most impressive accomplishments. The following list of parts is not intended to be
comprehensive, but rather gives an idea as to the variety of parts where worthwhile gains in performance have
been obtained:

1. Aircraft—Propellers, engine parts, wheels.


2. Marine—Propeller shafts, engine crankshafts.
3. Automotive—Coil and leaf springs, torsion bars, front axle spindles, crankshafts, wheels.
4. Railroad—Car axles.

Significant increases in performance have been obtained by prestressing techniques. Under certain
conditions, the fatigue strength of specimens has been more than doubled (3). 1 Improvements of this order
are not possible in all fatigue situations, nor in all materials. Furthermore, where they are possible, the
optimum prestressing conditions must be worked out by objective performance tests. Once these are
established and production specifications are set up, the processing engineer must insist that these
specifications be met on every piece processed. Mechanical prestressing is unique because there are no
simple, nondestructive, or even destructive, tests which can be routinely used to check whether a particular
part has been properly prestressed, although certain laboratory methods can be used to measure residual
stress. One must inspect the process rather than the part. Even then, subsequent processing steps can
reduce or cancel the benefits obtained. These steps could include honing or other finishing, straightening,
overheating, and so forth.

1. Numbers in parentheses in text and tables refer to references at end of report.

-5-
SAE J811 Revised AUG81

3.2 Improving Fatigue Resistance—An overwhelming majority of prestressing applications are aimed at
improving performance under fatigue loading conditions (cyclic stressing). A detailed discussion of the relative
contributions of residual stresses, strain hardening, and surface smoothness to fatigue resistance is beyond
the scope and intent of this report. Although it is admittedly oversimplified, acceptance of the premises that
initial fatigue cracking is associated with stresses in surface layers and that fatigue cracks are propagated by
tensile mean stresses will lead to at least a qualitative understanding of the role of prestressing.

The fatigue situations in which prestressing might be considered may be divided into four arbitrary classes as
follows:

1. "Normal" Conditions—No high stress gradients, no surface degradation from processing or service.
2. "Designed" Stress Concentrations—Fillets, grooves, transverse holes, and so on.
3. Surface Degradation From Service and Environments—Corrosion, fretting, wear, mechanical abuse
(nicks, gouges, and related misuses).
4. Negative Effects of Fabricating Processes—Machining, grinding, unfavorable heat treatment, plating,
anodizing, straightening, and so forth.

The data in Table 1 is illustrative of all these classes. Rolling increased the strength of the smooth specimen by
21%. The harmful effect of a sharp notch was cancelled when the specimens were rolled before notching.
Exposure to a corrosive medium before testing reduced the strength of a machined specimen to 69% of the
baseline. Companion specimens, which were rolled before exposure, not only exceeded the base line, but
were equal to those which were rolled, stored, and tested in laboratory atmosphere. An overbending (intended
to simulate a straightening operation), in which the smooth specimen was statically loaded in the test machine
to a total strain of 2.5%, left a very undesirable situation on the side of the specimen which was in compression
under the static load. The fatigue strength was reduced to 55% of the baseline, a more harmful effect than was
caused by the sharp notch. Although the strength of a similarly bent specimen was raised 73% by rolling, it
was not brought back to the baseline.

TABLE 1—RELATIVE IMPROVEMENTS IN FATIGUE STRENGTH OF A HIGH STRENGTH


ALUMINUM ALLOY PROVIDED BY A PARTICULAR SURFACE ROLLING, 1.5 IN. DIA.
SPECIMENS IN REVERSED BENDING (1)
1,000,000 Cycle Improvement
Treatment Strength, % of Due to Rolling,
Specimen “As Machined” %
As Machined 100 —
As Machined + Roll 121 21
As Machined + Notch 61 —
(0.020 in. deep)
Roll + Notch 105 71
(0.020 in. deep)
As Machined + Corrosion 69 —
Roll + Corrosion 122 77
As Machined + Overbent 55 —
Overbent + Roll 95 73

Table 2 shows how the bending fatigue strengths of steel crankshafts and specimens with fillets were
increased by rolling. Increases of 19% for the mild steel specimens and of 25% for the alloy steel crankshaft
were noted. A computed maximum contact pressure of 400,000 psi was used in rolling the specimens of mild
steel.

-6-
SAE J811 Revised AUG81

TABLE 2—IMPROVEMENTS IN FATIGUE STRENGTH OF STEEL:


(A) SPECIMEN OF 1.5 IN DIA WITH 3/8 IN FILLETS AND
(B) CRANKSHAFT FOR 4 1/4 × 5 IN DIESEL WITH 0.145 IN FILLETS.
EACH TESTED IN REVERSED BENDING (2)
Max Fillet Stress (ksi) Max Fillet Stress (ksi)
For 1,000,000 Cycle Life For 1,000,000 Cycle Life Improvement
Results From Steel Not Rolled Rolled Due to Rolling, %
Specimen (A) UNS No. G10430
SAE/AISI 1043
(Normalized) 42.0 50.0 19

Crankshaft (B) UNS No. G41400


SAE/AISI 4140
(300–340 Bhn) 55.0 69.0 25

Another case history is illustrated by Table 3. Mild steel shafts were first tested as normal type (rotating
bending) fatigue specimens, where it was established that 20% higher strength was obtained by rolling.
Specimens were then tested while held in a bushing where fretting was undoubtedly present at the maximum
stress area. The effect was to drop the strength of the ground specimen to less than half of those tested in
more conventional fashion. The fatigue strength in this environment was doubled by rolling. This almost
brought the strength back to that of the original ground specimen, tested without bushing.

TABLE 3—ENDURANCE LIMIT OF NO. 45 STEEL


(0.48% C, TENSILE STRENGTH = 92,500 PSI) (3)
Fatigue Strength Improvement
Test Specimen Limit (ksi) Over Ground Specimen, %
Conventional Rotating Beam:
Ground 39.0 —
Rolled 45.7 20

0.69 in dia Cylindrical, Tested


as Cantilever Beam Held in
Steel Bushing, Causing Fretting
Corrosion:
Ground 18.0 —
Rolled 36.2 101
Rolled + Light Grind 37.2 106

A comprehensive study of surface rolling was made by Atkin and Mezoff (4). It is concerned with the effect of
rolling on magnesium alloys and the development of a prestressed magnesium wheel for military use. Some of
the results are summarized in Table 4. Markedly different degrees of improvements were indicated for the
three alloys tested, ranging from 8–77%. No obvious pattern is established relating the relative degree of
improvements from rolling when loaded under complete reversal and zero-to-tension.

-7-
SAE J811 Revised AUG81

TABLE 4—FATIGUE STRENGTH OF MAGNESIUM ALLOY TEST SPECIMENS,


1.0 IN DIA WITH 5/16 IN RADIUS:
TESTED IN REVERSED AND UNI-DIRECTIONAL BENDING (4)
1,000,000 Cycle
Fatigue Strength 1,000,000 Cycle Improvement
Magnesium Alloy Magnesium Alloy As Fatigue Strength Due to
UNS No. SAE No. Loading Machined Rolled Rolling, %
M16600 (ZK60A) 524-F Reversal ±18.7 ±22.7 22
(Wrought) O-Tension 25.4 27.5 8

M11800 (AZ80A) 523-F Reversal — ±25.7 —


(Wrought) O-Tension 21.0 30.5 45

M11810 (AZ81A) 505-T4 Reversal ±11.5 ±19.0 65


(Cast) O-Tension 13.0 23.0 77

In this same reference, the improvement in fatigue life of actual wheels is also reported, as shown in Table 5. It
is characteristic of this product and test method that fretting and other surface damage may occur at high-
stress regions. This test showed that the life of a particular wheel design could be extended more than ten
times by surface rolling. Because of prestress due to tire inflation, the stress ratio in the critical regions would
be positive, i.e., from one level of tension to another. The specific amounts of improvement quoted are only
intended as representative figures. The actual amount possible in a given situation can vary widely in
response to many effects including specimen type, stress level, rolling techniques, and others.

TABLE 5—LIFE OF 20 × 7.50 MAGNESIUM WHEEL, CAST SAE 505-T4 [UNS M11810 (AZ81A)];
UNDER ROLL TEST WITH INFLATED TIRE AT 8,000 LB LOAD (4)
Wheel Description Life Cycles (Single Specimen) Location of Fracture
Not Rolled 326,000 Bead seat radius of
fixed flange

Bead Seat Rolled 2,018,000 Gutter cracks, bead


seat OK

Bead Seat and 3,946,000 Wheel OK—Test halted


Gutter Rolled

In summation, these examples show that fatigue strength can be improved by important—but widely varying—
degrees with prestressing methods. While worthwhile increases are observed on smooth or mildly filleted
specimens, the most impressive accomplishments are in protection against rather superficial surface damage,
typified by corrosion, fretting, or shallow notches. This type of damage can be surprisingly effective in reducing
fatigue strength of engineering structures where the surface is not prestressed.

3.3 Other Users for Prestressing—Stress corrosion cracking is a mode of failure that is influenced by surface
stress, environment, and time. A prerequisite to the start and growth of these cracks is the presence of tensile
stresses on the surface. Stress corrosion has been observed in many metals, including copper, lead, stainless
steel, iron, brass, magnesium, aluminum, and zinc. It has usually been associated with residual or assembly
stresses. Since it is strictly a surface phenomenon in its early stages and is entirely dependent upon the
presence of tensile stresses of considerable magnitude, considerable protection is offered by appropriate
prestressing. The major factor in this protection is certainly the introduction of compressive residual surface
stress.

-8-
SAE J811 Revised AUG81

There are uses of prestressing techniques for reasons other than, or supplemental to, strength improvement.
These, however, are usually so specific in terms of interest and material that they will be only briefly described
here. The following list is typical of some of these possible uses:

3.3.1 SURFACE HARDENING—When materials subject to work hardening are mechanically prestressed, the surface
layers may be significantly hardened. This may be useful in enhancing resistance to wear or other surface
damage.

3.3.2 SURFACE FINISHING—The "characterizing" of the surface obtained when prestressed is often a useful
engineering property. Surface rolling frequently produces a surface so smooth that subsequent grinding or
other finishing steps may not be needed. Ball burnishing and some of the techniques used on holes
(bearingizing, ballizing, taper plugs, and so on) can produce quite smooth surfaces.

3.3.3 CONTOUR FORMING—An example of this is the use of rollers or racks to produce strong, accurate threads and
splines. Another use of rollers is to form fillets to a desired contour.

3.3.4 SIZING—Some of the methods, under certain conditions, may perform a useful sizing function. Bearingizing
and ballizing are examples of this use.

3.3.5 SURFACE QUALITY CHECK—All the methods for mechanical prestressing depend on the yield strength of the
metal being exceeded, usually by a high, localized pressure. Certain types of metallurgical flaws at, or near,
the surface may be rendered obvious because of local spalling when subjected to this high stress. Thus, in a
limited sense, the prestressing also acts as a proof test of the surface layers.

3.4 Choosing a Prestressing Process—There are many situations where either shot peening or rolling may be
feasible to prestress and a choice is to be made between them. In this situation, there are several factors to be
considered:

3.4.1 GEOMETRY OF AREA TO BE PRESTRESSED—There are almost no limitations on the shape of areas that may be
shot peened, with the exception of small deep holes or slots. Rolling, on the other hand, can usually be
applied only to surfaces of revolution.

3.4.2 RESIDUAL STRESS DISTRIBUTION—The depth of residual compression that can be set up by rolling is
considerably greater than that set up by shot peening, although the surface compressive stress appears to
be about the same for both. While there are little available data comparing fatigue strength of identical
specimens prestressed by these two methods, one would expect that the potential for improvement by rolling
would equal, and probably exceed, that of shot peening.

3.4.3 CHARACTER OF SURFACE FINISH—Shot peening will produce a surface whose roughness is generally
dependent on the material hardness, shot size, and peening intensity. Rolling generally produces a much
smoother surface than shot peening.

3.4.4 COST—There are very little available data to support specific statements here. On large or complex areas,
shot peening will usually be cheaper, while on smaller areas of simple shape, rolling may be more favorable.
A factor that enters the cost situation is the large capital investment required for an efficient shot peening
machine, while rolling can be done in common machine tools with a modest tooling investment, unless highly
automated equipment is desirable. Parts often require masking before peening and cleaning afterwards.
Regardless of the comparison of these two methods, neither are expensive when compared to other
methods of precision finishing.

3.5 Limitations and Precautions—There are examples available showing that significant benefits are possible
through prestressing practices. As with all potent techniques, certain ground rules must be observed and
basic limitations recognized, or trouble may be encountered. In this section, some of these features will be
discussed briefly.

-9-
SAE J811 Revised AUG81

3.6 Introduction of Prestress—Perhaps the most obvious precaution here is that all the sensitive area of the part
must be covered by prestressing. The operation should cover enough area for its boundaries to be in
moderately stressed zones, and all of the intended area must be properly covered.

The size of the part needs to be considered in specifying rolling conditions. The volume of metal which has not
been plastically deformed by the working must be at least several times as great as that which has been
plastically deformed. This is necessary to insure that appropriate surface compression stresses are introduced
without unduly high internal balancing stresses of a tensile nature.

Since the introduction of prestresses depends on a nonuniform plastic action, in general, the dimensions of a
part will change when prestressed. In addition to this stress effect, there may be an apparent change in size as
a result of change in the geometrical character of the surface. This would be a "lowering" effect when a rough
surface is smoothed by rolling, or a "raising" effect when a smooth surface is peened with large shot or a
peening tool. Generally, none of these effects are large, but if fits or alignments must be precise, they should
be considered.

One of the more subtle limitations on mechanical prestressing is that there is no convenient and practical
method for inspecting the work produced. One must concentrate on control of the process. If this is not done
diligently, there is the risk of more or less prestressing intensity than intended. If on the low side, the
improvement that was counted on may not be obtained. If on the high side, some materials actually may be
harmed, with or without apparent visual indication of this damage.

3.6.1 MAINTENANCE OF PRESTRESSES—Once prestresses have been properly introduced, they obviously must be
kept present during service. Subsequent processing or unusual service conditions must not cancel the
benefits. Machining may cut away some of, or all, the compressive layer. Sufficient exposure to elevated
temperature may relax residual stresses. Straightening or stretching operations may reduce or neutralize the
deliberately introduced prestresses. Analogous effects may be present in service environments which may
jeopardize the effect that prestressing should have on performance.

3.6.2 FUNDAMENTAL LIMITATIONS—One must recognize that the improvements in some particular mechanical
property of a part which may be obtained through prestressing may lead to a change in some other property.
The compressive surface prestresses are invariably balanced by tensile stresses below the surface. This
internally balanced stress system may result in a lowered elastic limit or creep resistance. Under certain
conditions, the fatigue resistance may actually be reduced. Examples of this could be small diameter rolled
parts tested under axial loading or a relatively thin wall tube rolled externally and tested with fluctuating
internal pressure.

The excellent protection against surface abuse under fatigue loading that prestressing gives obviously is
limited in the sense that abuse exceeding some depth will be just as harmful, whether the part was
prestressed or not. The relationship of this depth to the intensity of working, and knowledge of whether there
is any notch depth where the part's fatigue resistance is actually harmed by prior prestressing, are points on
which little evidence is available, but which should be useful fields of research.

There are other properties where, in general, little or no improvement is expected. Static tensile strength is
one of these. Since this is the origin of the fatigue curve, the improvements should be related in some
increasing proportion with increased cycles. Part rigidity or stability are generally considered insensitive to
prestressing, as is impact strength. The latter property, however, may bear investigating for sensitive
materials.

3.7 Specific Characteristics of Various Materials—In the illustrations used earlier in this chapter, an attempt
was made to show responses of a variety of metals. Mechanical prestressing methods (other than shot
peening) have been applied to the following materials, with at least some degree of success: wrought steel, at
hardness levels ranging up to Rockwell C 45; gray, malleable, and nodular cast iron; cast and wrought
aluminum; and cast and wrought magnesium. This list is not intended to be exclusive of other materials.

-10-
SAE J811 Revised AUG81

Optimum prestressing intensities and specific peculiarities of these materials are matters on which little or no
published information is available. Certain materials show that a rather well defined optimum prestressing
intensity leads to maximum improvement. Others have shown the same order of improvement over widely
varying intensities. Some have shown good results with surfaces which were considerably degraded by the
prestressing operation, based on visual observation. Others have shown an actual loss in fatigue strength with
surfaces which, under casual observation, had an excellent appearance. Since all these observations were
based on specific, individual tests, it seems unwise to attempt to generalize the results. As stated earlier, the
choice of a prestressing method and intensity should be based on objective tests of the part under
consideration, or previous experience with similar parts under similar load conditions.

4. Theory of Surface Rolling for Fatigue Improvement (by C. W. Cable)

4.1 Introduction—Proper application of controlled surface rolling is an effective method for improvement of fatigue
performance. This chapter will discuss the reasons for this improvement and note some necessary conditions
associated with the typical residual stress systems obtained from surface rolling. For simplicity, the discussion
of residual stress distribution is confined to a two-dimensional presentation of unidirectional stresses. The
actual application of this theory to a particular part would necessarily include a stress analysis suitable and
adequate for that part. It will also be shown that, while local plastic deformation must occur during rolling to
obtain necessary prestressing, the combined residual and applied stresses obtained in service should be
within the elastic region; any yielding will change the residual stress system.

4.2 Effects of Surface Rolling—There are three principal changes caused by surface rolling to the part:

1. The surface finish is obviously "rolled," "roller burnished," and so on.


2. Surface layers of material are strain hardened or cold worked.
3. Surface layers are plastically deformed during rolling with the result that residual stresses are
produced.

Each of these effects can influence fatigue performance. A great deal of work has been done to investigate the
relationship of these factors especially with shot peening. The reader is referred to the references for a more
complete review of data on this subject. 2

It has been commonly accepted for many years that a smoother surface finish will result in fatigue improvement
due to the reduced extent of minute surface imperfections. Also commonly accepted is the fact that fatigue
failures usually originate at vulnerable stressed surface areas. Accordingly, we may expect that the finer
surface finish produced by rolling will contribute to fatigue improvement.

In considering the effects of cold working the surface layers on fatigue performance, it has been shown that,
within limits, both tensile strength and fatigue limit are increased by increasing amounts of cold work. The
fatigue limit, however, does not increase in proportion to the increase in tensile strength, i.e.: the endurance
ratio is reduced. Excessive cold working can even result in a decrease in fatigue strength (5). Since surface
rolling actually changes the surface layers of material by cold working, we may expect this to affect the fatigue
performance of the part.

These effects are generally considered to be secondary in importance to the effects of residual compressive
stresses. It has been shown that heating of shot peened surfaces will reduce fatigue performance, even
though hardness and surface finish were not significantly affected (6). Sigwart has reported fatigue
improvement by residual surface compressive stresses obtained by quenching (7). Horger has noted that
fatigue cracks were initiated at a rolled surface, but that the cracks then progressed at a slower rate than in
comparable specimens which were not rolled (11).

2. Numbers in parentheses in text and tables refer to references at end of report.

-11-
SAE J811 Revised AUG81

4.3 Residual Stresses—For the purposes of this discussion, only the simple case of uniaxial stress in a two-
dimensional section is illustrated. The necessary extension of the principles outlined herein to more complex
analysis can be carried out consistently with the stress analysis required for the particular part. The discussion
also assumes that we are dealing with a piece of homogeneous material, i.e., microstresses are not
considered here.

Consider a round bar section longitudinally as shown in Figure 1. It has a known or assumed residual stress
distribution near the surface as indicated by the heavy line showing the stress distribution across the section.
By application of two basic principles of statics requiring equilibrium of forces and of moments, for any
transverse section through Figure 1, it can be stated that:

1. The sum of the axial tensile and compressive forces must be equal to zero. If Section A-A, Figure 1,
were taken from a flat plate, the areas bounded by the curve and the zero stress line on the
compression side would equal the area on the tension side.
2. The opposing moments must be equal so that the sum of these moments equals zero.

The stress distribution curve is shown completed by the light line in a typical shape conforming to these
principles. Note that these do not define the intensity of the unknown stresses or their distribution; the exact
shape of the curve is not yet determined. It does provide insight, however, into the problems and also accents
a very significant point: whenever there is a given prestressed region, as at a surface, there must also be a
balancing region of opposite residual stress, usually of different intensity.

Figure 2 shows the same bar as in Figure 1 except that it has been bored out. Note that the tensile stress has
been increased, also the compressive stress has been reduced. Further, careful measurement will show that
the bar is longer.

Figure 3 shows a narrow strip cut from the tubular section of Figure 2. The strip is bowed, similar to the familiar
Almen strip used in shot peening. The compressive stress has been further reduced, but a compressive stress
at the opposite surface has also been produced with this relaxation. The necessary moment and force
equilibrium has been maintained.

In other words, prestressing means that the part has been mechanically processed so that it contains a desired
residual stress distribution which must be balanced both as to forces and moments. The exact distribution is
related to process and material.

FIGURE 1—LONGITUDINALLY SECTIONED ROUND BAR

-12-
SAE J811 Revised AUG81

FIGURE 2—LONGITUDINALLY SECTIONED ROUND BAR, BORED OUT

FIGURE 3—STRIP CUT FROM LONGITUDINALLY SECTIONED ROUND BAR, BORED OUT

The previous paragraphs have assumed a part with an existing desired residual stress distribution. To illustrate
how the desired stress distribution was accomplished, assume there is a piece of uniform homogeneous
material completely free from residual stresses. When, by definition, any loads are applied which produce only
elastic stresses, the part will return to its original condition upon release of these loads. Further, if subjected to
an assumed uniform plastic deformation, the shape will be altered, but no residual stresses will be produced.
Therefore, to produce residual stresses by mechanical means in a part consisting of one piece of a uniform
homogeneous material, nonuniform plastic deformation is necessary. This may be accomplished by locally
deforming the surface by rolling, or other processes.

The most significant point is that previous local nonuniform plastic deformation has produced the present
condition of static residual stress equilibrium. When known or determined, these stresses may be treated by
proper application of the theory of elasticity. Local plastic deformation has been illustrated by Sigwart, using,
as one example, the displacement and yielding of material under a Brinell ball (7). Essentially, high local
contact stresses produce local biaxial tensile yielding at the surface of the part when the deforming load is
applied. When this load is released, the yielded material will have been effectively stretched or spread out.
The normal complete recovery of the adjacent unyielded material from its elastic deformation is then
prevented. An equilibrium condition results where the yielded material is restrained in a biaxial residual
compressive stress state by the unyielded material which is essentially in a residual tensile stress state.

In the preceding section it has been stated that a desirable residual stress distribution would improve fatigue
performance. How this occurs will now be considered.

Figure 4A again shows a typical assumed residual stress distribution with surface layers in residual
compression. It has been shown that mechanical prestressing is obtained by plastic deformation of a local
area of initially uniform material and that the theory of elasticity must be applicable to the prestressed part. If
an external load is then applied to the part to obtain a given uniform load stress as in axial tension or a
nonuniform stress as in bending the applied load stresses and the residual stresses are additive algebraically.
This rule holds true as long as the stresses remain below the elastic limit at all locations; when unloaded the
part will elastically return to its original condition with residual stresses unchanged.

Superimposing an applied load tensile stress of less than the residual compressive stress allows the material
to remain in compression; Figure 4B shows a slight increase so that an applied tensile stress greater than the
residual compressive stress will cause the surface layer to be in tension, as in Figure 4C. The core material will
have a final tensile stress greater than the applied tensile stress. Where applied load compressive stress is
superimposed on an existing residual compressive stress layer the resultant compressive stress is increased,
as in Figure 4D. The allowable applied compressive stress, therefore, may be reduced, even though the yield
strength of the strain hardened surface layers will usually be significantly increased by rolling. Figure 4E shows

-13-
SAE J811 Revised AUG81

the effect of a residual stress distribution upon stress range when exposed to cyclic tensile and compressive
applied load stresses. The vulnerable surface layers under these conditions are exposed only to cyclic
compressive stress.

The effects of geometric stress concentrations such as grooves, oil holes, fillets, and so on, must be
considered both in compression and in tension to arrive at the resultant stress at any location. If the material
yield strength is exceeded by the resultant stress, the residual stress will be changed after release of the
applied load. For example: if an applied load compressive stress equal to 0.3 of the yield strength is added to
a local residual stress of 0.5 of the yield strength in the presence of a stress concentration with a factor of 2.5,
the calculated local resultant stress in compression equals 0.5 + (0.3 × 2.5) or 1.25 the yield strength.
Obviously, local yielding must occur. Note that the actual local residual stress intensity value of 0.5 yield
strength was used in this calculation and that this may be greater than the average residual stress at the
surface away from the stress concentration depending upon the relative size of the stress concentration and
the extent of the residual stress field.

Dugdale, in recent work, has shown some significant effects on fatigue performance obtained by prestressing
machined notched bars. Compression preloading resulted in a reduction of fatigue performance (8). This may
be explained as follows: it can be shown that a sufficiently large compression load applied to a notched bar
would result in yielding at the notch. Upon release of the load, the root of the notch would be in residual
tension. Such a residual tension stress at the notch would cause a reduction in fatigue life.

FIGURE 4—EFFECTS OF RESIDUAL STRESS DISTRIBUTION

-14-
SAE J811 Revised AUG81

4.4 Fatigue—In fatigue loading one is concerned with the effects of variable loads of less than the critical static
loads on structural integrity, part performance, or service life. The actual loads may be uniform alternating
loads or may be applied in a random manner and sequence. It is most important that any experimental or
analytical techniques simulate or duplicate actual service conditions as closely as practical, if reliable design
decisions are to be based on the results of using these techniques. Actual service is still the best proof of a
satisfactory design. For most experimental work a uniform, essentially sinusoidal, alternating load is used.
The mean stress may be zero, compression, or tension. The rotating beam test gives a zero mean stress.

A local surface residual compressive stress will affect only the resultant local mean stress; obviously it can
have no effect on the amplitude of the alternating stress. Any improvement in fatigue performance due to
residual stress, therefore, must come from a reduction of the mean tension stress at the surface. A Goodman
diagram, Figure 5A (9), has often been used to illustrate in a general manner the relative effects of mean and
alternating stresses on the fatigue limit. Note that the reduction of the mean tensile stress allows an increase
in the alternating stress, and, therefore, with a given alternating stress, reduction of the mean stress will
improve fatigue performance. Experimental data has indicated that this general relationship is true; however,
the exact linearity indicated by the Goodman diagram is not obtained. Figure 5B (9) shows a modified
Goodman diagram with a plot of typical fatigue life data. Figure 5C (9) illustrates another method used for
plotting such data. It is seen that a reduction of the mean stress should increase the fatigue life. The reader is
referred to the work of Grover, Dolan, and others for a more thorough review (9).

FIGURE 5A—GOODMAN DIAGRAM (9)

-15-
SAE J811 Revised AUG81

FIGURE 5B—MODIFIED GOODMAN DIAGRAM (9)

FIGURE 5C—EFFECT OF COMBINED AND STEADY STRESS ON FATIGUE STRENGTH

-16-
SAE J811 Revised AUG81

For the purposes of illustration in this discussion, assume a fatigue loaded specimen with a known surface
defect or crack of significant depth as shown in Figure 6. It is easily seen with this idealized defect that the
specimen can transmit compressive stresses across the crack without difficulty. A tensile stress, however, will
cause the crack to act as an extremely effective stress concentration. Plastic deformation will occur at the tip of
the crack, with the crack necessarily extending deeper into the material; repeated cyclic loading will continue
this damage until a critical cracksize is reached and failure occurs. By reducing the local surface tensile stress
in the vicinity of the defect, the damage per cycle will be reduced and longer fatigue life will be obtained.

4.5 Limitations of Theory—The previous sections have assumed that the material used was capable of sufficient
local plastic deformation during the rolling process to produce an adequately rolled surface without deleterious
effects on the material. It was also assumed that elastic behavior was obtained at stress level less than the
yield strength of the material. Further assumptions have been made that the material was originally free from
residual stresses and that thermal or other processes have not been introduced in the system.

If the ductile capacity of the material is questionable so that there is the possibility of surface defects being
initiated, a metallurgical examination is definitely indicated. Even when this condition exists, some
improvement may be possible if the depth of the deepest defect is less than the depth of the maximum
compressive stress; however, the danger here lies in the possibility of defects extending beyond the
compressive layer so that rapid failure could result.

Yield strength is an arbitrary definition of a measurable amount of nonelastic behavior for many materials, such
as 0.2% offset on a load-strain curve. The elastic behavior and yield strength of strain hardened material will
be improved over the original material. It follows that the foregoing references to yield strength should be
considered in this light with due allowance for the actual elastic behavior of the original and of the strain
hardened material.

FIGURE 6—EFFECTS OF VARIOUS TYPES OF STRESS UPON KNOWN SURFACE DEFECT

-17-
SAE J811 Revised AUG81

The existence of residual stresses prior to surface rolling complicates the picture, particularly if these are
present at a greater depth than the surface rolled layer. Surface rolling, however, can improve such parts,
especially where a thin surface layer is in residual tension.

It should be noted that elevated temperature exposure may result in creep or relaxation of residual stresses or
loss of residual stress by lowering of the yield strength. Similarly, severe temperature gradients and differential
expansion may alter or relieve residual stresses. In other words, subsequent processing and service
conditions should be examined to insure that the prestressed condition will be retained.

The actual maximum intensity of the residual compressive stress will vary with processing controls, but, for
heavy rolling of ductile material, it may be expected to slightly exceed the yield strength of the original material.
Residual stresses of this magnitude have been reported for shot peening by Brodrick (10), and Horger has
noted residual compressive stresses of over 100,000 psi longitudinal and 46,000 psi tangential from surface
rolling of normalized and tempered SAE 1050 steel (11). Far more information is available on fatigue
improvement than actual residual stress data.

4.6 Conclusion—Surface rolling may be expected to improve fatigue performance when properly applied to
suitable parts and materials. The process improves the surface finish, work hardens the surface layers, and
imposes a residual compressive stress on these surface layers. Of these three effects, fatigue improvement is
attributed principally to the super-imposition of a residual stress system upon the expected applied load stress
system. Thus, the local mean tensile stress at the vulnerable surface of the part is either reduced or reversed
so that the mean stress is actually in compression.

5. Calculations for the Surface Rolling Process (by H. R. Neifert)

5.1 Introduction—Successful application of the surface rolling process requires that the load on the roller be of
sufficient magnitude to cause plastic deformation in the surface material of the work piece. Plastic deformation
is obtained when the stress produced by the roller exceeds the elastic limit of the material being rolled. The
depth beneath the surface to which the rolling is effective is dependent upon the geometrical shape of the roller
and work piece, the materials involved, and the roller load. These factors can be related by mathematical
analyses which permit a determination of required roller loading based either on the maximum compression in
the contact area or on the shearing stress beneath the contact area.

The mathematical analysis assumes that the stresses produced in the contacting bodies are within the elastic
range of the material. Experience, however, has indicated that the theory may be extended into the plastic
stress range with reasonable accuracy for the practical application discussed herein.

LIST OF SYMBOLS

P compressive force between the two contacting bodies, lb.


Xx, Yy, Zz principal stresses in X, Y, Z directions, psi.
Sc maximum surface pressure, psi.
Ss maximum shearing stress (shearing stress in plane of maximum shear), psi.
Ss1 yield point of shaft material in shear, psi.
E Young's modulus, psi.
µ Poisson's ratio.
K material factor, psi.
Z depth below surface of contact measured along road centerline (Z axis), in.
a major semi-axis of the ellipse of contact, in.
b minor semi-axis of the ellipse of contact, in.
r1, r′1 radii of curvature at point of contact body No. 1, in.
r2, r′2 radii of curvature at point of contact body, No. 2, in.
δ form factor describing geometry of the two contacting bodies, in.
α, β constants describing the conformity of the contacting bodies.

-18-
SAE J811 Revised AUG81

φ angle between normal planes containing radii of curvature, r1 and r2.


Cos θ a parameter which always has a plus value.
P1 roller load for a desired depth of cold working Z1, lb.
m load factor = P1/P.
b B
e eccentricity of contact ellipse, --- or ----
a α
PMPc hardness, Rockwell "C" scale

5.2 Mathematical Analysis for Determination of Stress in the Contact Area—When two bodies in contact are
forced together by an external force, as in Figure 7A, stresses are induced in both bodies beneath the area of
contact. The three principal stresses, Xx, Yy, Zz (Figure 7B) are maximum at the contacting surface and
decrease as Z, the depth beneath the surface, increases. The problem of evaluating the stresses in the
surface of the contacting bodies was solved by H. Hertz (12, 13) 3. This work was later extended (14, 15, 16)
so that stresses beneath the area of contact could be calculated. The magnitude of the subsurface stresses
along the load centerline for a typical contact area is shown in Figure 8. The use of the subsurface shearing
stress in the computation of roller load will be discussed later in this chapter.

The maximum pressure between the two bodies, which acts at the center of the elliptical contact area, can be
calculated from the equation:

3 P
S c = --- ---------- (Eq. 1)
2 π ab
as indicated in Figure 7B.

It must be emphasized that the semi-axes of the pressure ellipse, a and b, are the projected semi-axes and are
not measured along the curvature of the pressure surface.

In the general case of compression of two bodies, where the surface of contact is an ellipse, the semi-axes are
given by the equations:

δ δ
a = α 3 P ---- ; b = β 3 P ---- (Eq. 2)
K K
The form factor, δ, describing the geometry of the two contacting bodies, and the material factor, K, are given
by the expressions:

4
δ = ------------------------------------------ (Eq. 3)
1 1 1 1
---- + ---- + ------ + ------
r1 r2 r 1 r 1
1 2

8 E 1 E2
K = --- ---------------------------------------------------------------- (Eq. 4)
3  2 2
E2 1 – µ  + E1  1 – µ 
 1  2

where E1,µ1 and E2,µ2 are elastic constants (Young's modulus and Poisson's ratio) of the materials of the two
bodies. The radii r1, r2, r'2 are the principal radii of curvature at the point of contact of the two bodies. The
radius of curvature is considered to be positive if the center of curvature lies within the body; if outside, it is
negative. (Example: the radius of curvature for a fillet on a shaft is negative.)

3. Numbers in parentheses in text and tables refer to references at end of report.

-19-
SAE J811 Revised AUG81

FIGURE 7A—BODIES IN CONTACT

FIGURE 7B—AN ELLIPTICAL AREA OF CONTACT OF TWO BODIES AND PRINCIAPL STRESSES
ACTING ON AN ELEMENT BENEATH THE CENTER OF CONTACT

-20-
SAE J811 Revised AUG81

FIGURE 8—MAGNITUDE OF PRINCIPAL STRESSES AND MAXIMUM


SHEARING STRESS ALONG LOAD CENTERLINE

In addition, planes A and B should be chosen so that plane A contains r1 and r2 and:

1 1 1 1
---- + ---- > ----- + -----
r 1 r 2 r' 1 r' 2

This is necessary to obtain positive values of Cos θ in Equations (5) and (6).

Plane A then determines the direction of the semi-minor axis of the pressure area and plane B the direction of
the semi-major axis of the pressure area.

The values of α and β are given in Figure 9 as a function of Cos θ where

δ 1 1 2 1 1 2 1⁄2
cos θ = ± ---  ---- – ----- +  ---- ± ----- + 2  ---- – -----  ---- – ----- cos 2 φ
1 1 1 1
(Eq. 5)
4  r1 r' 1  r 2 r' 2  r 1 r' 1  r 2 r' 2

and where θ is an angle between normal planes containing radii of curvature r1 and r2. Cos θ is merely a
parameter and the plus value is always used.

The angle θ will usually be 0 deg when performing a surface rolling operation, since the axes of rotation of both
bodies are generally in the same plane. Cos 2 θ is then equal to 1 and Equation (5) may be written in the
simplified form.

-21-
SAE J811 Revised AUG81

FIGURE 9—CURVES SHOWING VARIATION OF ΑΒ AND Ε WITH COS Θ


δ 1 1 1 1
cos θ = ---  ---- – ----- + ---- – ----- (Eq. 6)
4  r1 r' 1 r 2 r' 2

P
The ratio of -------------3- can be written from Equation (1) as follows:
( Sc )

3 2
P ( αβ ) δ
-------------3- = -----------------------------------
3
- (Eq. 7)
( Sc )  --------
1.5
K
2
 π 

5.3 Determination of Roller Load Using Known Sc—Roller loads for surface rolling of parts and materials for
which Sc has previously been determined may be computed using Equations (3), (4), (6), and (7). These
equations are usually used in applications of surface rolling where the depth of working desired is only in the
order of several thousandths of an inch, similar to that produced by shot peening.

The computation procedure may be briefly outlined as follows:

1. Calculate δ from Equation (3).


2. Calculate K from Equation (4). Values of K for several materials are given in Table 6.
3. Calculate Cosine θ from Equation (5) or (6) and read value of (αβ)3 from Figure 10.
4. Determine required or desired Sc by test, or use values recommended in Table 7 or Figure 11.
5. Put these values into Equation (7), and solve for roller load P.

-22-
SAE J811 Revised AUG81

TABLE 6—ELASTIC CONSTANTS AND MATERIAL FACTORS


FOR VARIOUS METALS ROLLED WITH STEEL ROLLER

3
Young’s Poisson’s  1.5
-------- K
2

Metal Modulus E, psi Ratio µ K, psi


 π 
Steel 29 × 106 0.25 41.2 × 106 1.85 × 1014
Aluminum 10 × 106 0.36 22.3 × 106 5.41 × 1013
Magnesium 6.4 × 10 6 0.28 15.3 × 10 6
2.55 × 1013

TABLE 7—VALUES OF SC FOR SEVERAL MATERIALS


(USED SUCCESSFULLY IN SPECIFIC CASES)
Material psi Source
Aluminum (Wrought) 362,000 G. A. Butz, Aluminum Company of America

Magnesium Alloy 258,000 R. L. Atkin and J. G. Mezoff (4)

Steel, Rc 30 400,000 J. O. Almen, also Figure 11 for steels of various hardness

TABLE 8—TABULATION OF α, β , e AND (αβ)3 FOR VARIOUS VALUES


OF θ AND COS θ (PLOTTED IN FIGURES 9 AND 10)

β
 ---
 α
θ Cos θ α β αβ)3
(αβ
0 1.0000 ∞ 0 0 ∞
10 0.9848 6.612 0.319 0.0483 9.383
20 0.9397 3.778 0.408 0.1080 3.662
30 0.8660 2.731 0.493 0.1805 2.441
35 0.8192 2.397 0.530 0.2211 2.050
40 0.7660 2.136 0.567 0.2655 1.776
45 0.7071 1.926 0.604 0.3136 1.574
50 0.6428 1.754 0.641 0.3655 1.421
55 0.5736 1.611 0.678 0.4209 1.303
60 0.5000 1.486 0.717 0.4825 1.210
65 0.4226 1.378 0.759 0.5508 1.144
70 0.3420 1.284 0.802 0.6246 1.092
75 0.2588 1.202 0.846 0.7038 1.052
80 0.1736 1.128 0.893 0.7917 1.022
85 0.0872 1.061 0.944 0.8897 1.005
90 0 1.000 1.000 1.0000 1.000

-23-
SAE J811 Revised AUG81

FIGURE 10—CURVES SHOWING VARIATION OF (ΑΒ)3 WITH COS Θ

FIGURE 11—SUGGESTED SC FOR STEEL OF VARIOUS HARDNESS (J. O. ALMEN)

-24-
SAE J811 Revised AUG81

For practical purposes, two assumptions are made in the application of these formulas. It will be noted that as
the roller moves up a fillet, the radius of curvature of that fillet is constantly increasing. However, this change in
radius, although numerically relatively great, has little effect on the value P/Sc3. Consequently, this slight
difference in value is assumed to be nonexistent. This same assumption is also applied to the roller when it is
brought in contact with the surface at some angle other than 90 deg.

Sample Calculation for Cylindrical Shaft (Example 1):

Shaft (Steel, Rc 30)


Diameter = 7.000 in = 2r′1
Contour Radius = ∞ = r1
Roller (Steel, Rc 62)
Diameter = 5.000 in = 2r′2
Contour Radius = 0.375 in = r2

Planes A and B have been chosen so that


1 1 1 1
---- + ---- > ----- + -----
r 1 r 2 r' 1 r' 2
and thus, Plane A contains the axis of rotation of the shaft and roller, andthe semi-minor axis of the pressure
area.
4
1. δ = ---------------------------------------------------------------------
1 1 1 1
---- + --------------- + --------------- + ---------------
∞ 0.375 3.500 2.500
4
= -------------------------------------------------------------------------------
0.666 + 2.666 + 0.286 + 0.400

= 1.193

2. K, for steel roller and steel shaft, where:


6
E 1 = E 2 = E = 29 × 10

µ 1 = µ 2 = µ = 0.25

4 E 6
= --- --------------2- = 41.2 × 10 , and
31 – µ

3
 15
------ K = 1.85 × 10
2 14
 π

1.193
3. cos θ = --------------- ( 0 – 0.286 + 2.666 – 0.400 )
4
= 0.592
3
( αβ ) = 1.325 (From Figure 10)

4. Sc = 400,000 psi (From Figure 11)

-25-
SAE J811 Revised AUG81

5. 2
1.325 ( 1.193 ) 3
P = -------------------------------------
14
( S c ) = 650lb
1.85 × 10
Sample Calculation for Shaft Fillet (Example 2):

Shaft (Steel, Rc 30)

Diameter = 7.000 in = 2r′1


Fillet Radius = 1.500 in = (−) r1

Roller (Steel, Rc = 62)

Diameter = 5.000 in = 2r′r2


Contour Radius = 0.375 in = r2

Planes A and B have been chosen so that


1 1 1 1
---- + ---- > ----- + -----
r 1 r 2 r' 1 r' 2

and again Plane A contains the axis of rotation of the shaft and roller, and the semi-minor axis of the pressure
ellipse.
4
1. δ = ------------------------------------------------------------------------------
1 1 1 1
– -------- + --------------- + --------------- + ---------------
1.5 0.375 3.500 2.500
4
= -----------------------------------------------------------------------------------
– 0.666 + 2.666 + 0.286 + 0.400

= 1.489

3
2.  15
------ K = 1.85 × 10 (Same as Example 1)
2 14
 π

1.489
3. cos θ = --------------- ( – 0.666 – 0.286 + 2.666 – 0.400 )
4
= 0.489
3
( αβ ) = 1.20 (From Figure 10)

4. Sc = 400,000 psi (Same as Example 1)


2
1.20 ( 1.489 ) 3
5. - ( S c ) = 920lb
P = ---------------------------------
14
1.85 × 10

-26-
SAE J811 Revised AUG81

5.4 Mathematical Analysis for Determining and Relating Subsurface Stresses—In the application of surface
rolling to large shafts and various heavy walled steel parts, the possibility of shape distortion is not as critical as
with parts of light section. Consequently, it has often been found desirable to surface roll heavy parts at roller
loads which will produce plastic working to a much greater depth than the superficial rolling performed on parts
of light section. Where such heavy pressures have been used on steel shafting, it is believed that the
improvement in fatigue resistance obtained is the maximum that can be developed for the material by the
surface rolling process. This contention pertains in particular to shafting having fitted members, such as
railroad axles, marine shafting, and so on. For example press-fitted assemblies, using a 91/2 in. dia SAE 1050
normalized and tempered shaft surface rolled to an effective depth of 1/2 in., have shown a resistance to
fatigue failure over twice that of assemblies not so rolled (17).

When surface rolling for deep penetration has been applied to such large parts, it has been found convenient to
determine the roller load by a different approach than previously described using the maximum contact
pressure Sc. The method used was suggested by S. Way (18) and relates roller load to the yield strength of
the material being rolled and the desired depth of penetration.

When the pressure area is circular, b = a, and the "eccentricity" of the ellipse of contact e = b/a = 1. For E1 =
E2 and µ1 = µ2, Z the principal stresses Zz, Xx, and Xy are given by the equations:
2
2aE a
Z z = --------------------------
2
- ------------------
2 2
(Eq. 8)
πδ ( 1 – µ ) a + Z

2
– ( 1 + µ )  1 – ---cot ---
2aE a Z –1 Z
X x = Y y = --------------------------
- ------------------ (Eq. 9)
2 2 2  a a
πδ ( 1 – µ ) a + Z

These are stresses along the load centerline (Z axis), and Z is the depth below the surface of contact in inches.
The maximum shearing stress at any point along the load centerline is equal to one-half the principal stress
difference, and is then found from:

1
S s = --- ( Z z – Y y ) (Eq. 10)
2

Substituting Equations (8) and (9) into Equation (10):


2
– 2 ( 1 + µ )  1 – ---cot ---
aE 3a Z –1 Z
S s = ------------------------------
- ------------------ (Eq. 11)
2 2 2  a a
2 πδ ( 1 – µ ) a + Z

Now if various values of depth below the surface are assigned to Z a shear stress curve as shown in Figure 12
is obtained. The maximum for this curve occurs at Z = 0.467a. Similarly, curves may be obtained for other
values of eccentricity "e". The family of curves plotted in Figure 12 can be used to permit development of a
semigraphical solution for determining roller loads.

-27-
SAE J811 Revised AUG81

FIGURE 12—MAXIMUM SHEAR STRESS AT VARIOUS DEPTHS ALONG LOAD CENTERLINE

The semi-axis of a circular contact area as obtained from Equation (2) is:

2
α = 1
3P δ ( 1 – µ )
a = 3 ------------------------------- 4E
4E K = -----------------------
2
-
3( 1 – µ )

For E = 29 × 106, µ = 0.25 and P = 1000 lb,

a = 0.02895 3 δ

For some other load P1 = PM = 1000 m,

a 1 = 0.02895 3 δ m, or a 1 = a 3 m (Eq. 12)

Z Z1
We may also set up the proportion, --- = 0.467 = ------
a a

Z1 Z
Then, ------------- = --- , and Z 1 = Z 3 m (Eq. 13)
a m a
3

Finally, to obtain the shear stress Ss for the Load P1, substitute into Equation (11) the values for a1 and Z1 from
Equations (12) and (13):

-28-
SAE J811 Revised AUG81

2
a3 mE 3( a3 m)
S s 1 = ------------------------------2 ------------------------------------------------
2
-
2
(Eq. 14)
2 πδ ( 1 – u ) ( a 3 m ) + ( Z 3 m )
 Z 3 m –1 Z m
3
– 2 ( 1 + µ )  1 + ------------- cot -------------
 a 3 m a3 m

Comparing Equations (11) and (14) we observe that:

Ss1 = S s 3 m (Eq. 15)

b
The relationships derived in Equations (13) and (15) are for circular contact where e = --- = 1. For elliptical
a
areas it can be similarly derived that:

Z1 = Z 3 δ m (Eq. 16)

m
S s 1 = S s 3 -----2 (Eq. 17)
δ

Eliminating m in Equations (16) and (17):


Z1Ss
Z = ------------- (Eq. 18)
δ Ss1

This equation represents a straight line through the origin of Figure 12.

5.5 Determination of Roller Load for a Specific Depth of Plastic Working—Roller loads for surface rolling to a
specific depth of penetration may now be computed by using Equations (16) and (18) and Figure 12. This
method is usually used in applications where the depth of plastic working required for maximum fatigue
resistance has been determined by previous testing, and is of some considerable magnitude.

The computation procedure may be briefly outlined as follows:

1. Compute δ from Equation (3).


2. Calculate Cos θ from Equation (5) or (6) and read the "eccentricity" e from Figure 9.
3. Determine desired depth of cold working from previous testing of particular material, or select from
Table 9.

-29-
SAE J811 Revised AUG81

TABLE 9—PRACTICAL MAXIMUM CALCULATED DEPTH OF COLD WORKING (Z1)


ATTAINABLE WITHOUT SURFACE DETERIORATION FOR MILD STEEL SHAFTING
(TENSILE YIELD POINT 40,000 TO 50,000 PSI)
Shaft Diameter, in Over 10 9.0 7.0 5.0
Z1, in 0.500 0.450 0.350 0.250

4. Compute Z from Equation (18), where

Z1 = desired depth of cold working, inch


Ss1 = yield point of shaft material in shear, psi = 1/2 (yield point in tension)
Ss = an assumed value of shear stress, psi

Plot Z, Ss on Figure 12 and draw a straight line through the origin. Where this line crosses the curve for the
"eccentricity" computed in Item 2, determine a value Z′.

5. Using Equation (16) solve for m:


Z1 3 1
m =  ------ ---
 Z'  δ

6. Required roller load, P1 = Pm = 100 m, lb

By this computation, a roller load will be selected which will produce, for a desired depth in the work piece,
shear stresses greater than the yield point in shear of the material. The shearing stress developed along the
load centerline will be as shown in Figure 13 and plastic yielding of the work material will take place to a depth
Z1.

This table is based on experience developed with 9 1/2 in dia press-fitted assemblies discussed by Horger
(17), where fatigue resistance to fracture within the press fit was increased over 100% by surface rolling. A
value of Z1 equal to 5% of the shaft diameter has been found to be a good rule of thumb for applying a similar
degree of surface rolling to mild steel shafting up to 10 in dia. For larger shaft sizes, surface deterioration limits
the depth of rolling to less than 5% of shaft diameter.

Unpublished information developed by Horger and Neifert on similar 5 3/16 in dia shaft assemblies shows that
surface rolling to a depth of 0.100 in will increase resistance to fracture by 50%. No curvature has been
established relating fatigue resistance of such large assemblies with depth of cold working up to the practical
maximum attainable. This example and others in the literature indicate that an appreciable increase in fatigue
resistance is obtainable with a penetration less than maximum.

Although resistance to fatigue fracture is greatly increased by surface rolling, the stress level at which fatigue
cracks initiate in press-fitted assemblies is little influenced. The effect of the compressive stress layer in
retarding the propagation of fatigue cracks has been shown by Horger (17). (see Figure 22.)

-30-
SAE J811 Revised AUG81

FIGURE 13—MAXIMUM SHEARING STRESS ALONG LOAD CENTER-LINE


COMPUTED FOR SHAFT AND ROLLER OF EXAMPLE 3

Sample Calculation for Cylindrical Shaft (Example 3)


Shaft (Steel, Rc 30)
Shaft (Steel, Y.P. = 50,000 psi)
Diameter = 9,500 in = 2r′1
Contour Radius = ∞ = r1

Roller (Steel, Rc = 62)


Diameter = 10.000 in = 2r′2
Contour Radius = 1.500 in = r2

Planes A and B have been chosen so that

1 1 1 1
---- + ---- > ----- + -----
r 1 r 2 r' 1 r' 2

and thus, Plane A contains the axis of rotation of the shaft and roller, and the semi-minor axis of the elliptical
contact area.
4
1. δ = ------------------------------------------------------------------
1 1 1 1
---- + --------------- + --------------- + -----------
∞ 1.500 4.750 5.00
4
= -------------------------------------------------------------------
0 + 0.666 + 0.211 + 0.200

= 3.71

-31-
SAE J811 Revised AUG81

2.
3.71
cos θ = ----------- ( 0 – 0.211 + 0.666 – 0.200 )
4
= 0.237

e = 0.74 ( from Figure 9 )

3. Z1 = 0.450 in (Determined by fatigue test (17))

0.450 ( 20, 000 )


4. Z = ------------------------------------------- = 0.0972in
3.71 ( 50, 000 ⁄ 2 )

5. Plot Z, Ss (0.0972; 20,000) on Figure 12, and draw a straight line through this point and the origin.
This line crosses the curve for e = 0.74 at Z1 = 0.105 in.

0.450 3 1
6. m =  --------------- ----------- = 21.3
0.105 3.71

7. P1 = 1000 m = 21,300 lb

If the maximum surface pressure, Sc, is computed for this example from Equation (1) a value of 650,000 psi is
obtained. This will emphasize the order of magnitude of surface pressure developed in the deep rolling of
heavy sections as compared with that used on parts of light section.

In surface rolling large shafts made from steel forgings having a yield point of 50,000 psi, experience has
indicated that deterioration of the shaft surface will occur if the surface stress exceeds a value slightly above
650,000 psi. This surface deterioration is evidenced by the appearance of metallic flakes in the lubricant, as
well as a scaly finish on the shaft surface. Such surface condition imposes a limitation on the magnitude of
surface stress that can be used on each particular material, and can be used as an experimental criterion for
determining maximum roller load in lieu of calculation.

5.6 Feed of Roller Along Shaft—Two factors must be considered in determining the optimum feed at which the
roller should traverse the work. The rate should be slow enough to permit uniform working of the surface, but
not so slow that the time consumed would make the surface rolling operation more costly than necessary.

As previously described, the contact area between the roller and work piece is generally an ellipse having
semi-axes a and b, Figure 7B. It has been found by test that a feed rate equal to the half-width of contact in the
plane of roller traverse will provide satisfactory coverage on steel shafts (19). The roller feed can therefore be
determined from the equations for the semi-axes of the ellipse of contact:
δ δ
a = α 3 P ---- ; b = β 3 P ----
K K
In computing roller feed, the necessity for proper selection of planes A and B should again be emphasized. If
planes A and B have been so chosen that:
1 1 1 1
---- + ---- > ----- + -----
r 1 r 2 r' 1 r' 2

-32-
SAE J811 Revised AUG81

plane A then determines the direction of the semi-minor axis of the pressure area and plane B the direction of
the semi-major axis of the pressure area. If plane A contains the axis of rotation of the work piece, the
maximum allowable feed will be:
δ
b = β 3 P ---- ( in ⁄ rev of work )
K
If plane B contains the axis of rotation of the work piece, the maximum allowable feed will be:
δ
a = α 3 P ---- ( in ⁄ rev of work )
K
The ratio:
Length of semi – axis in plane of roller traverse (in)
-------------------------------------------------------------------------------------------------------------------------------------- = 1
Roller feed (in)
may not be the optimum feed rate for all materials. Developmental work performed in connection with surface
rolling of magnesium alloy wheels indicated that maximum improvement in fatigue strength was obtained when
the above ratio was 10:1 (4).

Numerical computation of the feed rate, in addition to assuring proper coverage, can be useful in estimating the
time required for a specific rolling operation. In production rolling of parts, the time factor could influence the
selection of roller geometry as well as load.

Sample Calculation for Cylindrical Shaft of Example 1:

Plane A contains the axis of rotation of the work piece, so the maximum allowable feed will be:
δ
b = β 3 P ----
K
For Cos θ = 0.592, β = 0.66 (from Figure 9)

Substituting values for P, δ and K previously determined in Example 1,

b = 0.666 3 650  -----------------------------


1.193 
 06
41.2 × 10
= 0.017 in/rev

The feed for rolling the fillet in Example 2 would be similarly calculated.

5.7 Experimental Verification of Depth of Cold Working—The mathematical analysis discussed in this chapter
for computation of surface and subsurface stresses is theoretically correct only for loads which produce
stresses within the elastic limits of the material of the contacting bodies. This fact has been recognized in
extending the use of the formulas to computation of roller loads where the stresses are within the plastic range.

In the surface rolling of large mild steel forgings where roller load has been computed for a specific depth of
working, the error between aim depth and actual depth obtained has not been unreasonable. This is shown by
a typical example in Figure 14 where a hardness exploration on a transverse section indicated evidence of cold
working to a depth of 0.500 in, whereas the aim depth was 0.450 in.

Thus, although these formulas are for stresses in the elastic range, they provide a convenient tool for relating
rolling conditions of an unknown geometry to one which is known.

-33-
SAE J811 Revised AUG81

6. Tooling for Surface Rolling (by A. E. Di Gregorio and W. J. Fuhrman)

6.1 Introduction—The design and application of rollers and related tooling for refining and prestressing the
surfaces of fillets, threads, and bearings or journals, thereby increasing fatigue life, will be reviewed in this
chapter.

This material has been gathered from many large companies (21, 22, 24, 28, 29)4 which have conducted
original development work, and should be of value to engineers who desire to initiate experimental programs to
determine the benefits that can be derived relative to their own products.

Quite notably, job shop and toolroom techniques in surface rolling have laid the groundwork for many of our
current production accomplishments in this field—and with virtually identical and satisfactory results. For these
reasons the basic and simpler rolling procedures, their history and advantages, are presented as an
introduction to the production type tooling and machines presently employed for rolling of external cylindrical
surfaces and fillets of shafts.

6.2 History—The principal application of surface rolling to increase fatigue qualities is on fillets and plain or
threaded peripheries of shafts subjected to torsional and/or bending loads. In surface rolling, a suitably shaped
hardened steel roller under a special load is forced into the fillet or against the periphery while the shaft is
rotating for a predetermined number of revolutions, sufficient to smoothen the surface by plastic flow (26). (See
Chapter 4 for a more complete discussion of the reasons for fatigue strengthening.)

4. Numbers in parentheses in text and tables refer to references at end of report.

-34-
SAE J811 Revised AUG81

FIGURE 14—HARDNESS GRADIENT AT SURFACE OF 9 1/2 IN DIA CRANKPIN FORGING

As early as 1929, the beneficial effect in the prestressing of machined parts (29) through surface working by
pressure rolling was realized by the railroad equipment industry in their manufacture of heavy, highly stressed
components such as axle shafts and crankpins. Since that time, because of the large increase in fatigue
resistance obtainable, surface rolling of critical sections has also been widely acclaimed and accepted for
production applications in the automotive field.

Experience through the years has indicated that the success of surface rolling is contingent upon the proper
design of rollers mainly, and, particularly if increased hardness and fatigue life is to be realized, upon roller
pressures sufficiently great to cold work the material to an adequate depth. For example, portions of heavy
railroad axles and crankpins are surface rolled under pressures that produce a cold-worked depth of 1/2 in or
more (17). Although such operations will be discussed in the text, our interests will be concentrated mainly
upon improving fatigue endurance through the superficial effects resulting from the compressive residual stress
and the refinement of either machining or grinding imperfections, or both.

6.3 Simple Roller Devices—Rolling the periphery of straight external cylindrical surfaces on an experimental or
toolroom basis is normally performed on a lathe or similar equipment. The rolling operation essentially
consists of forcing a hardened steel roller under a predetermined load, applied by gravity, spring, air, or fluid
pressure against the revolving part and applying a transverse feed to effect a sufficient overlap of the roller
track, thus producing a cold-worked condition and a leveling off of tool marks to produce an improved bearing
surface.

-35-
SAE J811 Revised AUG81

(See Chapter 3). Determined by such factors as roller shape, pressure, feed, lubricant, and mechanical
condition of equipment, almost any degree of surface finish may be obtained by surface rolling. Although, in
most cases, fine surface finishes are required to resist wear on bearings and journals, the fatigue resistance of
rolled members is generally not dependent upon the degree of smoothness of the resulting surface (17). This
is particularly true in deep-worked rolling conducted by the railroad industry. In contrast to this phenomenon,
fatigue resistance of turned, polished, ground, or other types of machine-finished parts is largely influenced by
both direction and irregularities of finish. In fact, polishing and fine finishes are not necessary from the
standpoint of fatigue strength if the surface is to be rolled. When rolling, it is more essential to obtain an
adequate cold working of the surface layers to the proper depth rather than to obtain a low microinch finish
(17). A brief description of several special roller fixtures that can be applied to experimental type surface rolling
on a lathe follows.

Figure 15 depicts a relatively simple device, wherein a precalibrated scale is employed to facilitate known and
identical roller forces upon a succession of parts (29). As denoted, the tool steel roller, hardened to Rockwell C
62–65, is contained in a suitable yoke which in turn is hinged to a frame. The integral steel spring between the
yoke and the frame permits flexibility and provides a method for gaging the roller pressure by means of a dial
indicator. The device is mounted in the lathe tool rest on the compound slide. The roller, which might have a
profile radius of from 1/32–3/16 in (27), is mounted on tapered roller bearings which are adjusted to minimum
clearance within the yoke to eliminate any possibility of lateral movement of the roller. Such motion would
affect the smoothness of the finished rolled surface.

The steel spring can be designed to permit a range of deflection during various roller loads. A roller load
range, possibly from 400–3500 lb, and corresponding intermediate deflection readings, may be precalculated
and stamped on the frame for immediate reference (29). Without calculating the required forces, applied roller
loads should be of a magnitude sufficient to merely cause a visible deformation of the surface. To typify the
forces involved, theoretical contact stress values calculated from Hertz equations (defined in Chapter 3) for
hard steel would approach 800,000 psi (27).

Illustrated in Figures 16 (27) and 17 (25) are additional spring-loaded arrangements that can be utilized for
simple, lathe type rolling operations conducive to single or multiple external grooves and contours. Roller
pressures in these designs are adjustable by varying the preload of the compression spring used in either
case.

The three-roller device depicted in Figure 18 (17) is typical of the heavy-duty equipment employed by the
railroad industry. Such a device provides a satisfactory mechanical arrangement for applying a known
pressure on cylindrical shafts. The reaction from the rolling pressure is taken through the closed frame
supporting the rollers and is not transferred to the lathe centers as would be done with a single roller. The
entire frame is permitted to float on the front and rear carriage supports on the lathe; the carriage feeds the
rolling device along the shaft as though thread were being cut. This device has been designed to roll 6–15 in
dia shafts by reversing and changing eccentric bushings on the spindles supporting the individual rollers. The
top of the cradle opens so that a shaft may be loaded between the lathe centers without disturbing the rolling
device. Two rollers have a relatively sharp contour radius of 11/2 in (each applying 18,500 lb while a third roller
has a 5 in contour radius (applied at 21,000 lb). The former gives depth of work hardening and the latter a
smooth finish (17).

When applied to 9 1/2 in dia shafts in the region where a press-fitted gear, pulley, or wheel is to be applied, the
endurance limit against breakage in the fitted part is at least doubled. Depending on surface finish before
rolling, a hardness penetration of about 7/8 in can be obtained with only 0.001-0.003 in reduction in shaft
diameter. The depth of hardness penetration is determined by roller shape, size, pressure, and yield point of
the shaft material and will, therefore, vary for each application (29).

-36-
SAE J811 Revised AUG81

FIGURE 15—PRECALIBRATED ROLLER DEVICE

FIGURE 16—TYPICAL FIXTURE FOR ROLLING FILLETS OR EXTERNAL CYLINDRICAL SURFACES

-37-
SAE J811 Revised AUG81

FIGURE 17—SPECIAL FIXTURE FOR ROLLING EXTERNAL GROOVES AND CONTOURS

FIGURE 18—ROLLER DEVICE FOR HEAVY DUTY RAILROAD EQUIPMENT

Turned threads may be rolled in the root area to induce compressive residual stress and to improve fatigue
strength. Small bolts and screws are rolled with a hand tool, Figure 19. The bolt is held with one hand and the
tool is revolved with the other. Larger members are rolled with a lathe tool, Figure 20. The tool shank is a load-
calibrated spring which is held in the tool post. Cross-sections of the roller and thread, Figure 21, indicate
roller and thread geometry: A with no load on the thread, B with a roller force of 40 lb on the thread, and C the
approximate areas of residual compressive stressed material at the thread roots. Resulting fatigue
improvement is shown in Figure 22 for the three series of connecting rod bolts used in aircraft engines (30).

-38-
SAE J811 Revised AUG81

FIGURE 19—HAND TOOL FOR SUPERFICIAL ROLLING OF


SCREW THREADS IN SMALL BOLTS AND SCREWS (30)

FIGURE 20—LATHE TOOL FOR SUPERFICIAL ROLLING OF COARSE SCREW


THREADS IN WHICH SHANK SERVES AS LOAD CONTROL SPRING

-39-
SAE J811 Revised AUG81

FIGURE 21—ROLLER FOR SUPERFICIALLY ROLLING SCREW


THREADS IS DESIGNED TO COMPRESSIVELY STRESS METAL AT
ROOT SECTION OF THREAD (30)

FIGURE 22—EFFECT OF SUPERFICIAL THREAD ROLLING ON


FATIGUE DURABILITY OF CONNECTING ROD BOLTS (17)

6.4 General Rolling Procedures—The various roller devices that can be mounted in the lathe tool support should
preferably be so positioned that the centerline of the roller contacts the centerline of the work on the same
horizontal plane. This permits maximum and constant spring forces to be applied directly to the rollers,
facilitates stress calculations, and minimizes the resultant loads on the fixture and machine. When rolling large
fillets and radial relief grooves, the roller is inclined to the work at an angle equal to approximately half the
angle between the tangency point and the maximum position of rolling as depicted in Figure 23. On cylindrical
or tapered surfaces the roller may be fed straight into the work (29).

-40-
SAE J811 Revised AUG81

FIGURE 23—TYPICAL ROLLER POSITIONS FOR ROLLING FILLETS


AND RELIEF GROOVES

When applying manually operated rolling devices, the speed of rotation of the work being rolled is dependent
mainly upon the ability of the operator to maintain the proper rolling pressure with the cross feed or compound
feed, whichever is used. This is especially the case when rolling irregular and contoured surfaces. For the
inexperienced operator, speeds no greater than 25 rpm are advised until the rolling technique is mastered. As
the operator becomes more familiar with the operation the speed may be increased. A maximum surface
speed of 100 fpm appears to be reasonably satisfactory for an experienced operator (29).

As the work is revolved, the rate at which the roller transverses the work surface must be sufficiently slow to
obtain a certain degree of uniformity of cold working. Laboratory fatigue tests and service experience indicate
that for proper cold working, the roller traverse feed should not exceed one-half the width of the contact area
produced by the loaded roller upon the work (29). (Chapter 3 gives details in calculating feeds.)

In surface rolling large fillets and radial grooves, experience has proven that it is much easier to maintain the
required roller pressure by rolling down the fillet toward the axis of rotation. The desired roller pressure is
applied at the point of maximum roller position Figure 23 before the shaft is rotated, making sure that all the
backlash in the machine has been taken up in the direction of feed to prevent rolling a groove at the starting
point. With the longitudinal feed engaged and the work revolving, the roller is fed inward down the radius while
maintaining identical pressures over the entire rolled area. For full radius relief grooves, rolling is initiated at
either of the outermost points and continued down and past the center of radius on the root diameter. An
identical procedure is followed on the reverse radius of the groove, again rolling past center to insure a
completely rolled root diameter (29).

Prior to the actual rolling operation, the part should be thoroughly cleaned and demagnetized if necessary to
remove any particles of steel that could be forced into the parent material by the rollers. Throughout the rolling
operation, a straight mineral or mineral-lard oil of low viscosity (or kerosene) should be used for lubrication.

-41-
SAE J811 Revised AUG81

6.5 Fillet Rolling—Due to the inherent design of many components, fillets at the juncture of two different
diameters are normally subjected to the greatest torsional and bending loads. Consequently, it is in this area
that premature failures occur. Lack of sufficient cross-sectional area or a drastic change in cross-sectional
area, however, are not the only deficiencies. In numerous instances, the surface finish of the fillet governs its
resistance to fatigue. This is true even where finish grinding is employed, and, in many cases, fatigue
resistance is definitely aggravated by such operations. In these cases, as the part is contacted, the dressed
radius on the grinding wheel breaks down. The typical fillet then produced on the part is not a true radius, (see
Figure 24) resulting in stress points that reduce fatigue strength. Failure to hold part print tolerance at the fillet
through grinding is common throughout the automotive industry, and is probably most serious on crankshaft
bearings. Rolling the critical fillets results in work-hardening of the surface to a depth of approximately 0.0001
in without the apparent tear marks produced by abrasive methods. An approximate 115% increase in fatigue
strength of nodular iron crankshafts is substantiated by actual engine and laboratory tests (21).

FIGURE 24—TYPICAL DEFICIENCIES OF GROUND FILLETS

Throughout years of experimentation to increase fatigue life of critical sections by surface rolling, probably the
earliest automotive application and positive results were achieved by fillet rolling the highly stressed areas on
the pin and main bearings of gray iron engine crankshafts. Here, cold rolling of fillets by means of steel balls
increased the limiting stress 60% for reverse bending and 80% for "one-way" loading as compared to machined
and ground fillets (20). In this method, as depicted in Figure 25 (20), three hardened steel balls are equally
spaced around the fillet and held in place by a split loading ring having an internal 45 deg chamfer. With two of
these rings assembled on the crank-pin, it is possible to roll both fillets at the same time, the ring being loaded
endwise by a simple device containing four coil springs compressed by a screw attachment. This device can
be easily rotated by hand during the actual rolling operation which is comprised of only several revolutions. To
produce a small but detectable deformation of the fillet, typical loads applied to the rings are approximately one
ton (20). A deformation of about 0.003 in in the fillet results, and, consequently, a work-hardened surface of
very smooth and truly circular form. A small ridge of metal is normally raised at the junction of fillet and journal
(possibly 0.0005 in) (20). To alleviate this condition, roller pressures should be maintained at a minimum and
allowances made in the mating product component to insure that this slight ridge does not become an
interference point (21). Although this simple method of fillet rolling can be quite successful, it often induces
deformation and failure of the relatively small balls. Because of the geometry, the balls afford no practical
solution for their retention and control, which is particularly essential for production applications.

As indicated, the success of surface rolling is mainly contingent on the design of the roller. This is true
particularly in rolling smaller fillets. Although relatively large fillets and relief grooves can be experimentally
hand followed on a lathe, as has been discussed, it is the smaller fillets (3/8 in and below) which primarily
concern the critical strength areas normally encountered in highly stressed crankshafts, front wheel spindles,
and the like.

-42-
SAE J811 Revised AUG81

FIGURE 25—BALL DEVICE TO ROLL CRANKSHAFT FILLETS

For rolling concentric fillets such as on main bearings of crankshafts, toolroom methods employing spring
loaded devices for use on lathes (see Figure 16) can likewise be applied to fillet rolling except that special
rollers are recommended; these being fed into the corner fillet at 45 deg or at a position bisecting the shoulder-
to-axis angle. No feed is necessary for these fillets.

The roller mounting and method of generating and maintaining the required force vary considerably with
different users. The roller is usually pressed lightly onto a pin which rotates in bushings mounted in the holder.
Sufficient side clearance is given to the roller so that it can slide laterally and center itself in the fillet (26).

Currently, two types of rollers are in commercial use for rolling small fillets.

Plain Fillet Roller—A plain type roller, as illustrated in Figure 26, (26), is used effectively for fillets having a
radius of 1/32 in or less. The roller periphery has a radius at the low limit of the fillet radius tolerance, which
generates, by plastic flow, a new shaft radius conforming to the roller periphery. An applied force of less than
100 lb is usually sufficient for this small radius, and greatly simplifies both the roller and shaft mounting. Very
few shaft revolutions are required to obtain the maximum effect; in one application 10 revolutions of the shaft
are sufficient. Plain rollers of oil-hardening tool steel machined to approximate the dimensions noted in Figure
26 and hardened to Rockwell C 62–65 have served for several thousand pieces at a negligible replacement
cost (26).

Wedge Roller—It would be most desirable if all fillet rolling could be accomplished with a plain roller by making
the radius on the roller periphery larger as the radius of the shaft fillet is increased. However, above 1/32 in
fillets, the required force on the roller would rapidly reach high values where the roller and shaft mountings as
well as the means of generating the required force become quite complex (26). (The normal manufacturing
variation in fillet radius is also an important part in this same problem.) Through extensive studies, the wedge
type roller (Figure 27), was developed (22). (C. W. Jackman, U.S. Patent #2,357,515.) This unique roller
design offers a vast improvement. It can roll large fillets with a relatively small applied force and at the same
time is unaffected by the normal manufacturing variations in fillet radii (26).

-43-
SAE J811 Revised AUG81

FIGURE 26—PLAIN FILLET ROLLER AND APPROXIMATE DIMENSIONS

B
For Fillet
Radius, in A, in deg min C, in D, in E, in F, in
1/16 to 3/32 0.174 1 57 0.030 1.625 0.437 0.031
3/32 to 5/32 0.284 3 58 0.060 1.625 0.437 0.031
5/32 to 1/4 0.435 6 04 0.080 1.562 0.460 0.0625
1/4 to 3/8 0.612 6 52 0.090 2.125 0.687 0.0625

FIGURE 27—WEDGE-SHAPE FILLET ROLLER

-44-
SAE J811 Revised AUG81

On the wedge-type roller, the wide portion of the wedge rolls the extremities of the fillet while the narrower area
of the wedge works the center, as shown in the two detail sketches in Figure 28 (26). The other points on the
wedge contact the intermediate points in the fillet. Because the various points of contact between roller and
fillet are virtually identical in area, the applied forces are equalized about the periphery of the roller. As the
shaft rotates, the center of the roller with respect to the work never changes, but the roller edges ride up and
down on each side of the fillet in a sort of kneading action (26). Since the relation of the diameter of the roller
and diameter of the surfaces on which it works is varied with respect to each other, each point of the roller
surface applies a different condition to each point of the fillet surface throughout the operation, and the entire
fillet surface is rolled completely and uniformly.

FIGURE 28—ACTION OF WEDGE-SHAPE ROLLER ON FILLET

The dimensions shown in Figure 27 are recommended for rolling shaft radii of different sizes. When selecting
or designing a roller diameter, it is important that the roller speed be nonsynchronous with the work speed so
as to distribute the work roller contact. It is desirable, therefore, to have the circumference of the roller such
that it is not a multiple of the circumference of the work. This suitable difference in circumferences will give a
hunting action to insure proper rolling of the fillet.

A modified version of the described wedge-shaped roller design has been developed (28). This concept not
only adds considerably to the quality of the rolled fillet, but provides a definite simplification for its manufacture
on toolroom equipment. On this type of roller, the narrowest section of the wedge is ground to a full radius
approximately 0.010 in under the low manufacturing limit of the fillet to be rolled. This full radius is carried
symmetrically on the periphery of the roller for approximately 90 deg as illustrated in Figure 29.

FIGURE 29—MODIFIED WEDGE ROLLER

-45-
SAE J811 Revised AUG81

As with the plain fillet roller, oil hardening tool steels at Rockwell C 62–65 have given excellent results. The
inherent offset of the modified wedge roller can be machined and ground by toolroom methods consistent with
the typical operational process illustrated in Figure 30. A special cam grinding machine highly expedites the
cam generations of the wedge-type roller (28).

FIGURE 30—TYPICAL PROCESS TO MANUFACTURE MODIFIED WEDGE-SHAPE ROLLER

Experience has proven that the speed of the work while the rollers are in contact is immaterial. More cycles
are required to roll a fillet with the wedge roller than with a plain roller. While using a single wedge roller, 200–
600 rotations of the part is common, however, the use of multiple rollers simultaneously engaged will reduce
the rotation requirements accordingly. Applied roller pressures of 60–200 lb (26), and a theoretical
compressive shear stress of approximately 650,000 psi (28) of contact are also typical.

Normally, at comparable fillet radii, life of the wedge roller is less than for plain rollers because of the longer
rolling time; however, wedge rollers have endured 15,000–20,000 pieces (26). Replacement cost is, therefore,
negligible.

A light oil is especially necessary for lubrication while fillet rolling with the wedge roller. And, as a general rule
for all types of surface rolling applications, rolling should not be combined with metal removing operations
which inherently induce loose particles of metal or abrasives into the rolled fillet.

-46-
SAE J811 Revised AUG81

6.6 Fillet Rolling Applications—A typical example of the advantageous use of fillet rolling to effect a substantial
increase in fatigue life can be illustrated by a specific problem that beset a large automobile manufacturer. In
the manufacture of front wheel spindles having a Brinell hardness of 302–341, extremely poor machinability
with respect to tool life and related factors was experienced. To resolve this situation the Brinell hardness was
reduced to 255–285; however, this would have likewise reduced fatigue life of the spindle at the critical inboard
bearing fillet. Here, a stress concentration existed due to the drastic change in diameters. Figure 31 is typical
of the resulting ultimate failures produced when subjecting this critical area to a bending moment in laboratory
tests (21).

FIGURE 31—TYPICAL ULTIMATE FAILURE INDUCED BY DRASTIC


CHANGE IN CROSS-SECTIONAL AREA

To maintain comparable fatigue endurance at the reduced hardness, the critical fillet areas of extruded front
spindles, finish ground to size, were surface rolled by hand in the experimental machine illustrated in Figure 32
(21). The experimental rolling fixture contained three rollers equally spaced around the stem of the part. The
lower two rollers were fixed while the upper roller was moved by means of a lever arm acting on a toggle
arrangement. Each roller had a 0.125 in radius at one end to coincide with the critical fillet of the spindle. This
radius projected 0.002 in from the roller proper so that initial contact between rollers and part was made at the
critical fillet. The spindle was rotated at 210 rpm by means of a floating driver on a lathe. A thrust screw
through the chuck was tightened prior to the rolling operation to force the part shoulder against the rollers to
insure a fully rolled fillet. Rolling time was approximately 20 s, during which time a light lubricating oil was
applied. The toggle was adjusted so that a pull on the lever of approximately 100 lb would cause the toggle to
reach dead center. The resultant pressure was dependent upon the resistance met by the rollers and was not
known.

Subsequently, the production type machine illustrated in Figure 33 (28) was installed for fillet rolling front wheel
spindles. Here, mounted to the machine base, is a motorized head, the spindle of which accepts a center and
drive plate to locate and rotate the part at 400 rpm. Opposing the head is a hydraulically operated tailstock to
support the shaft end of the part. Between the head and tailstock are two hydraulically operated slides
positioned horizontally 45 deg to the center of the part; one forward and one rearward of the rotational axis.
Each cylinder, at 400 psi, feeds and withdraws a wedge-type roller to the fillet. Semiautomatic controls actuate
the tailstock center, initiate rotation of the spindle, and simultaneously feed the two fillet rollers to the work.
After a predetermined rolling cycle, which in this case is 15–17 s, the rollers retract and the machine stops for
unloading. A light oil is used for lubrication during the rolling operation. Replacement of rollers is required only
after 15,000–20,000 pieces. A production rate of 127 pieces/h is maintained.

-47-
SAE J811 Revised AUG81

FIGURE 32—EXPERIMENTAL DEVICE USED FOR FILLET ROLLING FRONT SPINDLES

FIGURE 33—PRODUCTION MACHINE FOR FILLET ROLLING FRONT SPINDLES

Where multiple fillets are involved or the rotation of the fillets is not concentric with the central axis of the shaft,
such as on the main and pin bearings of crankshafts, a unique tooling problem is encountered (26).

As a toolroom method for accomplishing such rolling of crankshaft fillets, the device pictured in Figure 34 has
been experimentally employed (24). This is comprised basically of a yoke containing two sets of two plain
rollers positioned to float in the radius of opposing bearing fillets. The yoke is mounted to the carriage of a
lathe in such a manner as to allow lateral float to compensate for the throw of the pin bearings. Opposing the
two sets of fillet rollers is a support roller which remains in parallel contact with the bearing to be rolled. It is
against this roller that the adjustable spring pressure is applied to force the rolers into the fillets. Use of this
design is inherently limited to the width of the bearings which are to be fillet rolled. Inadequate width would not
allow sufficient space for the two rollers. Because of the variable direction in which the rollers are presented to
the fillet, plain rollers in place of the preferred wedge-type rollers become a necessity.

-48-
SAE J811 Revised AUG81

FIGURE 34—EXPERIMENTAL DEVICE TO ROLL CRANKSHAFT FILLETS

To provide an experimental means of fillet rolling with wedge rollers or toolroom type equipment which would
eventually serve as a basis for adaptation to production, the toggle-type arrangement depicted in Figure 35
was devised (28). It is mounted to a crossarm attached to the carriage of a lathe and actuated through a 2.7:1
mechanical advantage by hydraulic pressure. This action scissors two opposing roller shoes against the main
and pin bearing diameters of the crankshaft. The inherent design permits the entire head to reciprocate
vertically to compensate for the throw of the crankpins. Being mounted to the carriage, the head is also
adjustable laterally to allow for flexibility with various crankshaft designs for which rollers and shoes are special.
Each of the opposing roller shoes carries two wedge rollers mounted at a 45 deg angle to the line of the
crankshaft bearings and so arranged that when the shoes are close around the crankshaft bearings, the four
rollers in each pair are diametrically opposed. All rollers are permitted to float laterally to compensate for
crankshaft bearing tolerances.

Roller pressures are applied by means of a hydraulic cylinder incorporated within the head, through connecting
rod and toggle linkage to claming fingers which support the roller shoes. This pressure is adjustable to suit
whatever conditions require, and remains constant after the correct amount of pressure is determined.

In order to achieve proper results from this operation, it is first necessary machine all fillets of a crankshaft to
the limits specified on the part print, not to exceed plus or minus 0.018. The wedge rollers are generated for a
specific size fillet, and any variation beyond these limits will defeat the defect of the operation (21).

-49-
SAE J811 Revised AUG81

FIGURE 35—ADJUSTABLE DEVICE FOR EXPERIMENTAL ROLLING


OF CRANKSHAFT FILLETS AND JOURNALS

Employing such an arrangement to roll fillets of 0.110–0.120 in radii, roll pressure required on main and pin
bearings previously machined to size is 324 lb. Each bearing fillet requires an approximate 30 s cycle or 75
revolutions of the part at 150 rpm (21).

Following the apparent success of the aforementioned process to experimentally fillet roll crankshafts, the
basic toggle-type head was incorporated into a production type machine (28). These heads are spaced in
series to accommodate each bearing position as illustrated in Figure 36. The machine consists of a fabricated
steel base with suitable driving mechanism, including hydraulic and coolant systems. Head and tailstock are
mounted on the top of the base. The tailstock is adjustable laterally to provide for loading of various crankshaft
lengths. A crossrail mounted on trunnions at the back of the machine is aranged for mounting the workheads,
the number of which depends on the number of main and pin bearings of the crankshaft to be worked. The
workheads reciprocate and adjust similar to the original experimental concept.

This machine is automatic, hydraulically operated with push-button control. The operator loads the part and
with the push button, lowers the workheads. Roller shoes automatically close on each crankshaft bearing.
After the operator pushes the button to start the machine, it operates for a predetermined time and
automatically stops. The cycle reverses for unloading and reloading. Actual pressures and speeds simulate
the experimental version. Time for the simultaneous rolling of all fillets is approximately 30 s.

6.7 Journal Rolling Crankshafts—As a means of eliminating the lapping operation of crankshaft bearing
diameters and its inherently high cost, while maintaining comparable surface finishes and at the same time
increasing fatigue strength and wear qualities through superficial hardening, rolling of the journals has proven
beneficial (21).

-50-
SAE J811 Revised AUG81

FIGURE 36—PRODUCTION MACHINE ADAPTABLE FOR FILLET OR


JOURNAL ROLLING OF CRANKSHAFTS

The same basic toggle-type equipment used for both experimental and production fillet rolling of crankshafts
can be employed for rolling the bearing surfaces, except that different rollers and shoes are necessarily
required. Here, helical concentric rolls depicted in Figure 37 are employed (28). These are mounted again in
special shoes so that, circumferentially, the roller helix is alternated between left- and right-hand lead. In
contact with the revolving crankshaft, the rollers laterally knead the high points of the surface into the lower
areas, thereby improving the finish of the ground surface from approximately 25-12 µin, rms. Figure 38
illustrates typical roller dimensions for journal rolling. To minimize the roller-to-work contact area and, thereby,
the required working forces, the diameter of the rolls should be as small as practicable. Applied forces,
revolutions, and time requirements remain identical to the basic fillet rolling machine (28).

Figure 39 indicates a variation to the toggle-type arrangement for journal rolling (23). Results here are
identically successful in rolling crankshaft journals. In this machine the crankshaft is loaded between centers
and on two fixed rolls to support two of the main bearings. Subsequently, the mechanical crank-operated
head, synchronized in rotation with the throw of the pin bearings, lowers and allows two opposing and
predeterminately spaced rollers at each position to straddle the bearing diameter as indicated in Figure 40.

6.8 Conclusion—The tooling, methods, and results as outlined in this chapter are representative of experimental
work and actual production applications. In each instance, however, the final results were governed by such
factors as roller design, product design, material, and material hardness. It remains, therefore, the
responsibility of the engineer to evaluate the data as it pertains to the specific problem at hand; each case is
peculiar to itself.

7. Control and Inspection (by D. J. Wulpi)

7.1 Introduction—The reliability of a part is usually dependent upon the necessity for making the production parts
at least as good as the preproduction parts that were tested in the engineering phase of development. Many
manufacturing processes lend themselves to fairly accurate inspection after processing; others make such
subtle changes in the part that inspection is difficult. Mechanical prestressing processes are in the latter
category, for the most potent strengthening factors—compressive residual surface stresses—are very elusive
and difficult to evaluate properly.

-51-
SAE J811 Revised AUG81

FIGURE 37—WORK HEAD TO SURFACE ROLL CRANKSHAFT JOURNALS

FIGURE 38—DIMENSIONS OF TYPICAL JOURNAL ROLLER

-52-
SAE J811 Revised AUG81

FIGURE 39—PRODUCTION MACHINE FOR JOURNAL ROLLING CRANKSHAFTS

FIGURE 40—PRINCIPAL OF IMPCO SURFACE ROLLING MACHINE

-53-
SAE J811 Revised AUG81

7.2 Control—Since it is difficult to inspect finish rolled parts, the most practical method for consistently obtaining
desired properties in production is to maintain close control of the rolling procedure and also to make frequent
fatigue tests of the finished parts. Although usually not critical, the roller load and geometry should be kept
fairly constant. With respect to the part, the hardness and prior surface finish may be expected to change the
fatigue strength if they are allowed to vary widely. When rolling metals where relatively little plastic deformation
is expected, the surface should be free from deep tool marks, so that the roller will contact the surface as
intended. If there are "steps" or deep tool marks in a surface, prior to rolling, they may well result in improperly
finished parts of relatively low fatigue strength. During the rolling process, the part and the rollers must be free
from foreign matter such as dirt, chips, and so on.

7.3 Surface Finish—Although there is little direct evidence, the surface finish on a part, prior to surface rolling,
apparently is not critical to fatigue strength, provided the entire surface in the zone subject to high service
fatigue stress is plastically deformed. For this reason, higher rolling pressures must be used with hard
materials and rough surfaces than are necessary for softer, smoother parts. It is believed that prior surface
finish is more critical when little plastic deformation is involved than when a greater degree of plastic
deformation takes place. Prior surface finish should be held constant for given rolling conditions if relatively
close final dimensional tolerances are necessary, as for a shaft in a press fit (56). 5

7.4 Nondestructive Inspection—As previously mentioned, control of the rolling process is important because of
the difficulty of inspection of the finished part. At the present time, apparently, there is no commercial device
for nondestructive inspection of rolled surfaces. Some experimental work has been done, primarily with X-rays
and electro-magnetic induction, or "eddy current" testers, but no devices are commercially available as yet. In
these tests, both of which detect changes due to plastic deformation, the X-ray instruments measure variations
in the crystallographic structure of the metal, while the electromagnetic induction, or "eddy current", devices
measure changes in electrical and magnetic properties of magnetic materials. It is hoped that commercial
instruments will become available for nondestructive inspection of rolled surfaces.

There are, however, some simple methods for assuring that a part has been surface rolled and, to a certain
extent, how well it has been rolled. A stripe of paint, ink, or similar material may be applied to the surface to be
rolled and to the adjacent surface (57). Inspection will readily indicate whether or not rolling has been
performed over the stripe because of the different appearance of the stripe in the rolled and nonrolled zones.
For critical parts which have previously been subjected to magnetic, dye penetrant, ultrasonic, or other critical
inspection controls, it is recommended that use of such controls be continued, with inspection both before and
after rolling. Once a reliable process control procedure has been established for the rolling operation and
performance reliability of the rolled parts confirmed, the necessity for the use of such critical inspection controls
can be reevaluated (58).

A properly rolled surface should have a smooth appearance, but may be either bright or dull in luster. No tool
marks from prior machining should be evident. Change in dimension may sometimes be used as a control, but
this is influenced to a large degree by the flattening of tool marks.

7.5 Destructive Inspection—The metallurgical effects of rolling can be detected in different ways with different
metals. If the effective rolling is sufficiently deep on a metal subject to work hardening, the increase in
hardness may be detected by microhardness traverse. Where the rolled zone is shallow, however,
metallurgical inspection is quite difficult. Figure 14 of Chapter 3 shows the hardness gradient formed by very
heavy pressure on 9 1/2 in dia normalized carbon steel shaft (56). The vertical line at the 0.450 in depth
indicates the depth of work hardening calculated by the method outlined in Chapter 3. Excellent correlation is
usually obtained when a microhardness traverse is used on steel to check this method of calculation (56).

5. Numbers in parentheses in text, tables, and figure titles refer to references at end of report.

-54-
SAE J811 Revised AUG81

Metallographic examination can frequently be used to measure the depth of cold working. In ferrous metals,
however, it is sometimes difficult to distinguish this deformed layer by metallographic means. In some
nonferrous metals the effect is readily observed: in aluminum by recrystallization following solution treatment,
Figure 41 (57); in magnesium by twinning within the grains, Figure 42 (4).

FIGURE 41—SAE AA 2014 AND UNS A92014-T6 ALUMINUM (WROUGHT)


AFTER SURFACE ROLLING. SPECIMENS REHEATED TO SOLUTION TEMPERATURE
AND QUENCHED IN COLD WATER. MAGNIFICATION: 45 × (57)

8. Effect of Surface Rolling on Properties Other Than Fatigue (by H. O. Fuchs)

8.1 Introduction—In surface rolling, a thin surface layer is forced to flow by compressing it—spot after spot—
under the roller. Many changes in the part are thus produced; the changes are all related to each other but will
here be considered under three separate headings:

a. Changes of the surface geometry (generally increased smoothness).


b. Changes of the material in the deformed layer (cold work).
c. Changes of the entire rolled part (residual stresses).

These changes, inherent in the rolling process, increase the fatigue strength as discussed in other chapters.
Their influence on other properties is discussed in this chapter.

-55-
SAE J811 Revised AUG81

FIGURE 42A—SAE 505 AND UNS M11810-T4 MAGNESIUM (CAST)


SHOWING STRUCTURE PRIOR TO SURFACE ROLLING.
MAGNIFICATION: 35 × (4)

-56-
SAE J811 Revised AUG81

FIGURE 42B—SAE 505 AND UNS M11810-T4 MAGNESIUM (CAST)


SHOWING HEAVILY TWINNED STRUCTURE NEAR SURFACE
AFTER ROLLING. MAGNIFICATION: 35 × (4)

-57-
SAE J811 Revised AUG81

a. Surface Geometry—Surface rolling has long been used to produce smooth surfaces on parts such as
journals. Table 10 shows some reported results. Horger (17)6 shows profilograph records and gives
numerous references to other such records.

TABLE 10—
Hardness Finish Finish
Before Rolling, Before After
Item Material Brinnel Rolling, µ in rolling, µ in Reference
1 Steel, AISI 195 120 6 (31)
B1112
2 Steel 290
20 to 35 12 to 16 (32)
Alloy No. 4 (141,000 UTS)
3 Stainless Steel
SAE 51416 210 90 8 (31)
UNS 514600
4 Cast Iron 185 200 15 (31)
5 Brass 116 80 10 (31)
6 Aluminum,
SAE AA2017-T4 102 150 5 (31)
UNS A92017

Item 2 was a tapered shaft, rolled with a roller of 1.25 in dia with a thread radius of 0.025 in under a
pressure of 71 lb, with a feed in the last of two passes of 0.0015 in/rev. The other items were holes
rolled with long small diameter rollers which are pressed against the surface hundreds of times per
second while revolving in the hole. (Chapter 7 gives a more complete discussion of this "bearingizing"
process.) The improved surface finish was accompanied by an increase in hole diameter of the order
of 0.0005 in. Reduction of 0.002 in by surface rolling a 9 1/2 in dia rolled steel shaft was reported (17).

b. The Cold Worked Layer—Many materials change in hardness when they are forced to flow. "Work
hardening" is often observed; work softening of previously work hardened material has also been
reported (33).

Table 11 lists hardness readings for the surface rolling hole specimens for which surface finish was
given in Table 10 and for shafts of 0.50% carbon steel with 96,000 psi tensile strength.
Other changes also occur in metals which have been forced to flow; grains change shape;
transformation or precipitation may be promoted; the modulus of elasticity of hard steel may
temporarily be decreased by cold deformation to about 90% of its original value; magnetic properties
may be improved; grains may coalesce on recrystallization (this may be used for destructive
inspection).

c. The Entire Rolled Part—Surface rolling sets up compressive stresses in the plastically deformed skin
which are balanced by tensile stresses in the core of parts. These stresses have a profound influence
on the behavior of the part. Their beneficial effect on fatigue is treated in other chapters of this manual
and will not be covered in this chapter.

6. Numbers in parentheses in text, tables, and figure titles refer to references at end of report.

-58-
SAE J811 Revised AUG81

TABLE 11—
Hardness Hardness Thickness of
Below Skin, at Surface, Hard Skin,
Material Brinell Brinell in Reference
Steel AISI B1112 195 240 0.015 (31)
Stainless Steel, SAE 51416 and UNS 541600 210 273 0.004 (31)
Cast Iron 185 260 — —
Brass 116 155 0.013 (31)
Aluminum, SAE AA2017-T4 and UNS A92017 102 122 0.005 (31)
Steel, 0.50% Carbon 170 220 0.500 (17)

If the depth of the compressed layer is excessive relative to the section size, the tensile stresses in the core
may reach high values and eventually lead to failure in the core of hollow parts. It should be understood that
this is a rare condition. In unusual parts the core has so much more section area than the skin that tensile
stresses remain quite low. In the absence of specific tests it is suggested that the compressively stressed skin
should have a cross section area of no more than 20% of the total cross section.7 This will hold the tensile
stress in the core down to about 10% of the peak compressive stress near the surface. (If more precise tests
are not available one might estimate that the depth of the compressed layer, or the thickness of the
compressed skin, is roughly equal to the width of the indentation made by the roller (34). (See also Chapter 5.)

In a tensile test the residual stresses will lead to yielding at a lower load stress because, in some areas of the
test piece, the load stress plus the residual stress exceeds the yield strength before the load stress alone
would exceed it.

The compressed skin of a rolled part will prevent or greatly retard the growth of cracks. Besides fatigue
resistance this also increases the resistance to stress corrosion greatly or prevents damage by stress
corrosion entirely. For instance, a steel shaft of 141,000 psi tensile strength rotating in brackish water with a
bending load producing 20,000 psi stress lasted 3 million cycles (2 days) without surface rolling, but with
surface rolling it lasted more than 100 million cycles (50 days) (32).

Surface compression also prevents the brittle fractures which result from many plating procedures. This effect
has been well documented for shot peening (6, 36), and is similar for surface rolling.

Similarly, the effect of surface compression in extending the range of ductile behavior of steel to lower
temperatures (37) may safely be expected if surface rolling is used as the means of producing the
compression.

The harmful effects of fretting in promoting fracture are also prevented or greatly delayed by surface rolling.
Increases in endurance strength under fretting conditions have been reported and are shown in Table 12.

Where pinhole porosity is a problem as in some parts which must be plated or must hold pressure, the surface
compression induced by surface rolling can be used to close surface pores and cracks.

TABLE 12—
Permissible Stress, psi Permissible Stress, psi Reference
Without Rolling With Rolling
20,000 60,000 (32)
11,000 22,000 (17)
800 3,200 (38)

7. Or a depth of no more than 5% of the diameter.

-59-
SAE J811 Revised AUG81

9. Other Methods for Mechanical Prestressing (by J. G. Roberts)

9.1 Introduction—Mechanical prestressing of structural members by shot peening (39) and surface rolling have
been described previously. Other methods for mechanical prestressing may be classified according to the
initial cause of the prestress, as shown below.

TABLE 13—
Yielding of Material
Surface Yielding
Presetting
Shock Hardening
Subsurface Yielding
Bearingizing
Ballizing
Hammer Peening
Strain Peening
Stress-Relief Grooving
Interference of Assembled Parts
Press Fitting
Miscellaneous
Stamping
Polishing

Other than shot peening and surface rooling.

Each of the methods herein, with the exception of press fitting, causes residual stress by yielding (40)8 of
material at or near the surface.9 Consequently, some material ductility is always required. In fact, the more
ductility the material possesses the better it is suited for prestressing. Virtually all structural engineering metals
and alloys have some ductility and may, therefore, be mechanically prestressed. In addition to generating
residual stresses, the localized plastic deformations cause some materials to strain harden. This, too, can
have a beneficial effect. Of course surface finish is altered by many of the prestressing methods. Depending
upon the initial surface finish and/or the prestressing method and its severity, the prestressing can smoothen,
roughen, or leave unaltered the initial surface finish. In describing the prestressing methods listed in Table 13
particular attention will be directed at the resulting residual stress, strain hardening, and surface finish. Some
of the examples used to describe these methods will illustrate only the strain hardening or some other
characteristic of the prestressing. Residual stress will, however, always be induced.

9.2 Presetting—One of the simplest methods for increasing fatigue strength of a part that will be service loaded in
only one direction is to statically stress the structural member in the same direction it is loaded in service to a
fixed maximum load or strain that exceeds the elastic limit. The fixed load or strain is usually held for a short
time. Upon release of the load the member will have a residual stress distribution which will permit the same
(initial) load to be subsequently carried elastically. As a result of the yielding, the member will also be
permanently deformed. Material strain hardening is not required.10

The presetting of leaf spring specimens is an illustration, Figure 43 (42). Here, one-directional bending
endurance limit is increased from 90,000 psi to 130,000 psi. All of this increase is attributable to the
compressive residual stress induced by presetting the SAE 5160 steel. Torsion bars (43) and Belleville springs
(44) furnish other examples of parts that have been preset.

8. Numbers in parentheses in text, tables, and figure titles refer to references at end of report.
9. For a general mathematical description of how residual stresses may occur as a result of plastic deformation, see (40) Initial Stresses, p. 425.
10. For a description of residual stresses produced by inelastic bending, i.e., presetting, see (41), p. 375

-60-
SAE J811 Revised AUG81

Rectangular cross section beams can be preset to a greater extent than I-beams.11 Residual stresses caused
by presetting circular shafts12 by one-directional torque and by presetting thick cylinders by internal pressure13
are given by Timoshenko.

9.3 Shock Hardening—High velocity stress waves14 can cause plastic deformations which induce residual
stresses; however, judging from the experience reported in the literature, it appears that it is used primarily on
materials that strain harden appreciably. The surface finish may be unaltered by the stress wave.

Railroad switch frogs, Figure 44, have been hardened by high velocity stress waves generated by sheet
explosive (45). Five times its former useful life (20 versus 4 years) is expected for this application.
Photomicrographs show plastic flow of the crystal structure that results in the hardness increase shown in
Figure 45 (45). The penetration of the hardened layer extends to about 1 in, which is considerably deeper
penetration than is usually found for other prestressing methods. Residual stress data were not reported, but it
would appear that a considerable portion of the increased life is due to the strain hardening of the surface layer.
Also, surface finish is apparently not altered greatly. Austenitic manganese steels, weldments, and castings
have also been hardened by this method (46).

Fatigue S-N curves of explosively formed materials indicate that the high velocity stress waves do not always
increase fatigue strength (47). Difficult-to-form aircraft materials such as 8% Mn-Ti have better fatigue strength
in the annealed condition than in the annealed and hardened (by shock hardening) condition.

9.4 Elastic Contact Stresses—Before describing the prestressing methods which are initially caused by
subsurface yielding (see Table 13) of the material it is instructive to briefly review the maximum stresses
caused by static loading of a ball on a flat plate. In Figure 46 15 the three normal stresses along the normal
(Oz) to the center of the circular contact area are shown. Distance below the contact circle is shown (vertical
scale) in terms of a, the radius of the circle of contact, and the stresses, which are compressive, are shown
(horizontal scale) in terms of the maximum compressive stress qo. The latter is usually referred to as the
"Hertz" stress. It occurs at the center of the contact area. The maximum shearing stress occurs at a depth of
0.5a and is equal to 0.31 qo, which is almost 2 1/3 times the maximum tensile stress (0.133 qo), which occurs
at the boundary of the circle of contact. Ductile steels first yield at the point of maximum shearing stress.
Since the depth at which the material first yields affects the penetration of the residual stress, it is useful to
express the radius of the contact circle in terms of qo and R, the radius of the sphere. For steel, a = 2.86 qoR/
E. Thus, the distance below the surface at which yielding first occurs varies directly with both qo and R, and
inversely with E, the modulus of elasticity in tension.

After considerable yielding has taken place, Figure 47 (48), the elastic stresses are altered by the plastic flow
and the computations are no longer valid. They do provide, however, a rational means for studying the
penetration and the amount of yielding.

11. See Timoshenko (41), p. 371


12. See Timoshenko (41), p. 383.
13. See Timoshenko (41), p. 389.
14. For a mathematical description of how such waves are generated, see (40), p. 438.
15. From (40), p. 376.

-61-
SAE J811 Revised AUG81

FIGURE 43—RESIDUAL STRESS INDUCED BY PRESETTING LEAF


SPRING SPECIMENS AND RESULTING FATIGUE PERFORMANCE (42)

-62-
SAE J811 Revised AUG81

FIGURE 44—SHOCK HARDENING OF RAILROAD-SWITCH FROGS (45)

FIGURE 45—HARDNESS INDUCED Y SHOCK HARDENING


RAILROAD-SWITCH FROGS (45)

-63-
SAE J811 Revised AUG81

FIGURE 46—COMPRESSIVE CONTACT STRESSES ALONG


PROJECTED LOAD AXIS FOR A BALL PRESSED AGAINST A FLAT PLATE (40)

FIGURE 47—CALCULATED PLASTIC REGION IN FLAT PLATE (48)

9.5 Bearingizing—Rapid blows of rollers against the bearing surface for the purpose of strain hardening the
material and/or surface finishing the bore is known as "bearingizing." A tool similar to that shown in Figure 48
(17, 49) is usually used for this purpose. The bores of the link rod shown in Figure 49 (17, 49) were rolled with
such a tool with the resulting fatigue improvements shown.

A cutaway view of a "through" tool is shown in Figure 50 (50). As high points on a cam come up behind a
series of rollers they deliver an outward radial blow to peen the work surface.

-64-
SAE J811 Revised AUG81

FIGURE 48—BEARINGIZING TOOL (17, 49)

FIGURE 49—FATIGUE IMPROVEMENT IN LINK ROD BY BEARINGIZING (11, 12)

-65-
SAE J811 Revised AUG81

FIGURE 50—BEARINGIZING "THROUGH" TOOL (50)

Overall views of a "contour" tool and a "through" tool are shown in Figure 51 (50).

FIGURE 51—BEARINGIZING TOOLS (50)

-66-
SAE J811 Revised AUG81

9.6 Ballizing—Through holes may be prestressed quickly and precisely by forcing an oversized ball through the
hole. This induces residual compressive stresses around the hole and reduces the vulnerability to fatigue
failure; however, the process is usually applied in production as an economical means for precision sizing and
finishing holes.

Fishbolt holes in railroad rails (51) have been prestressed by this method with a 50% fatigue strength increase
which is maintained after exposure to corrosion. Selection (52) of the approximate ball size and determination
of the final size, finish, roundness, surface hardness, and taper of the finished hole depend upon the following
variables:

1. Material.
a. Hardness.
b. Ductility.
c. Heat treatment history.
d. Cold working history.
e. Strain hardening properties.
2. Hole geometry before ballizing.
a. Out-of-round.
b. Inside diameter tolerance.
c. Surface roughness.
d. Waviness of inside diameter.
e. Wall thickness.
3. Ballizing process.
a. Method of supporting part.
b. Lubricant used.
c. Interference fit between ball and hole.

The experimental method is usually used in arriving at the desired characteristics of the finished hole.

9.7 Hammer Peening—Large parts may be conveniently and inexpensively prestressed by peening with a large
radius tool (frequently 1 in). In this way greater penetration or depth of the prestressed layer may be obtained.
Large gear teeth, Figure 52, have been prestressed in this manner to offset fatigue failure (53). An air hammer
tool is fitted with a spherical-tipped hammer made of hardened (Rockwell C 60 +) tool steel. Air pressure
drives the hammer into the work for a distance determined by the porting of the driving piston. Usual air
pressures are in the range from 35–70 psi and resulting impact rate ranges from 35–70 impacts/s. Hammer tip
radius is determined by depth of penetration desired, part shape, and material ductility. For a given air
pressure, small tip radii tend to increase depth of penetration.

9.8 Strain Peening16

Structural members that are shot peened while held at some fixed strain are said to be strain peened.
Unidirectionally stressed members may be strain peened to increase fatigue strength.

Leaf spring specimens of SAE 5160 steel and 50 Rc hardness have been strain peened (42). The initially flat
specimens are shot peened17 on the face loaded in tension during fatigue testing (test surface) while held to a
fixed longitudinal curvature. After peening, the "free" curvature will vary, depending upon the amount and sign
of the applied strain during peening. Likewise, longitudinal residual stress will vary, depending upon the
amount and sign of the applied strain during peening, and longitudinal residual stress will vary systematically,
Figure 53, and will greatly influence the fatigue limit. The S-N curves of Figure 53 indicate that the fatigue limit
of conventionally shot peened specimens may be increased by about 1/3 by strain peening.

16. See U.S. Patent No. 2,608,752.


17. Chilled iron shot, 0.006 C Almen intensity, full visual coverage.

-67-
SAE J811 Revised AUG81

FIGURE 52—AIR HAMMER PEENING LARGE GEARS. FAILURE


AVERTED BY HAMMER PEENING IN PLACE. MINE HOIST GEARS
12 FT DIA × 30 IN WIDE (53)

FIGURE 53—RESIDUAL STRESS AND FATIGUE PERFORMANCE OF


STRAIN PEENED LEAF SPRING SPECIMENS (42)

-68-
SAE J811 Revised AUG81

9.9 Stress-Relief Grooving (17, 61)—Fatigue cracking around transverse oil holes of crank pins may be offset or
eliminated by stamping stress-relief grooves around the hole, Figure 54 (17). Steel specimens (SAE 1050 and
UNS G10500) with a yield strength of 50,200 psi and a tensile strength of 92,000 psi and having the
dimensions shown in Figure 54 were stamped with the apparatus shown. The rotating bending endurance limit
was increased from 15,500 psi for the shaft without grooves to over 20,000 psi with stamped grooves, or an
increase of over 35%. Coined grooves around holes may also be used to strengthen the holes in fatigue (35).

FIGURE 54—STRESS-RELIEF GROOVING

9.10 Press Fitting—Press fits and shrink fits are well known methods for holding tube assemblies together. Such
interference fits also give rise to elastic stresses which may be beneficial. This is illustrated by an example
(41). If in Figure 55, the radii a, b, c are 2, 4, and 6 in, respectively, and the interference is 0.005 in., the
tangential residual stress distribution in the steel cylinders is shown by the line mm in the inner cylinder and by
the line m1n1 in the outer cylinder. An internal pressure of 30,000 psi will cause a stress distribution in a solid
cylinder shown by the line SS, and in the built up cylinder, the superimposed stresses are shown by the
ordinates to the shaded area. Thus, the maximum tangential stress is reduced from 50,000 psi to 42,000 psi
and is moved from the inner radius of the interface radius.

This method for mechanical prestressing does not require plastic deformation.

FIGURE 55—PRESS FIT ASSEMBLY CROSS SECTION (A) AND (B)


TANGENTIAL STRESSES (41)

-69-
SAE J811 Revised AUG81

9.11 Stamping—This method for prestressing involves cold stamping the edges of holes, Figure 56 (17), Burnheim
(54) found that steel plates of the shape shown in Figure 56 could be stamped at the hole edges to increase
the reversed bending fatigue limit up to 44% by using the conical pin stamp and loading shown in the lower
right portion of Figure 56. The conical stamp and loadings shown at the lower left resulting in fatigue limit
improvements of from 25–33%. The Cr Mn V steel plates (yield strength 88,200 psi) with the hole unstamped
had a 34,100 psi fatigue limit.

FIGURE 56—TOOLS FOR STAMPING HOLE EDGES (17, 54)

9.12 Polishing—Mechanically polished surfaces can cause fatigue strength improvements by reducing the stress
concentrations and by prestressing the surface. Coombs' (55) reversed bending fatigue tests of shot peened
leaf-spring steel indicate that hand polishing to a 7 µin finish causes about 20% increase in fatigue limit.

PREPARED BY THE SAE IRON AND STEEL TECHNICAL COMMITTEE

-70-
SAE J811 Revised AUG81

Rationale—Not applicable.

Relationship of SAE Standard to ISO Standard—Not applicable.

Application—The two most widely used methods of mechanical prestressing probably are surface rolling and
shot peening. Since the process of shot peening has been rather widely discussed in previously
published literature, the greater part of this manual is concerned with surface rolling and its theory, load
specification, tooling, control, and effects. Methods briefly considered include hammer peening, cold
pressing, and treatment of small holes with balls or tapered pins. The general aims of this manual are:

a. To give the reader a general understanding as to what mechanical prestressing is and whether it
may be expected to help him with his product.
b. To help him choose a process.
c. To help him get started in tool design and preparation of test samples. At this stage of development
of the art, the optimum prestressing conditions and degree of performance improvement should be
established by objective and destructive tests, unless one has ample previous experience on similar
materials and products.

Reference Section

Butz, G. A., and Lyst, J. O., "Improvements in Fatigue Resistance of Aluminum Alloys by Mechanical
Surface Prestressing." Paper presented at 1961 Western Metals Congress (ASTM).

Gadd, C. W., Anderson, J. O., and Martin, D., "Some Factors Affecting the Fatigue Strength of Steel
Members," SAE Transactions, Vol. 63, 1955.

Kudryavtsev, I. V., "The Influence of Internal Stresses on the Fatigue Endurance of Steel," Proceedings
of the International Conference on Fatigue of Metals, Institution of Mechanical Engineers, London;
ASME. New York, 1956.

Atkin, R. L., and Mezoff, J. G., "Development and Testing of Magnesium Alloy Wheels." Paper presented
at Third Sagamore Ordnance Materials Research Conference, December 1956, Syracuse University
Research Institute.

Lessels, J.M., Strength and Resistance of Metals, New York: John Wiley and Sons, Inc., 1954.

Cohen, B., "Effect of Shot Peening Prior to Chromium Plating on the Fatigue Strength of High Strength
Steel," WADC Technical Note 57-178, ASTIA Document #AD 130821, 1975.

Sigwart, H., "Influence of Residual Stresses on the Fatigue Limit," Proceedings of the International
Conference on Fatigue of Metals, Institution of Mechanical Engineers, London; ASME. New York,
1956.

Dugdale, D. S., "Effect of Residual Stress on Fatigue Strength," The Welding Journal, January 1959.

Grover, H. J., Gordon, S. A., and Jackson, L. P., Fatigue of Metals and Structures, Prepared for Bureau of
Aeronautics, Department of the Navy, Nav Aer OO-25-534, 1954.

Brodrick, R. F., "Protective Shot Peening of Propellers—Residual Peening Stresses," WADC Technical
Report 55–56, Part I, June 1955.

Horger, O. J., "Cold Working," Section 6.9, ASME Handbook, Metals Engineering—Design, McGraw-Hill,
1953.
SAE J811 Revised AUG81

Hertz, H., Journal of Mathematics (Crelles' Journal), Vol. 92, 1881.

Hertz, H., Gesammelte Werke, Vol. 1, p. 155, Leipzig, 1895.

Belajef, N. M., "On the Problem of Contact Stresses," Bulletin, Institute of Engineers of Ways and
Communication, St. Petersburg, 1917. Memoirs on Theory of Structures, St. Petersburg, 1924.

Thomas, H. R., and Hoersch, V. A., "Stresses Due to the Pressure of One Elastic Solid Upon Another,"
University of Illinois Experimental Station Bulletin No. 212, Vol. 27, No. 46, July 15, 1930.

Lundberg, G., and Odqvist, F. K. G., Proc. Ingeniors Vetenskapa Akad., No. 116, Stockholm, 1932.

Horger, O. J., "Stressing Axles and Other Railroad Equipment by Cold Rolling," Surface Stressing of
Metals, American Society for Metals, pp. 85–142, Cleveland, Ohio, 1947.

Way, S., Discussion of paper by R. E. Peterson and A. M. Wahl in Journal of Applied Mechanics, Vol. 2,
No. 2, June, 1935, pp. A-69–71.

Horger, O. J., "Effect of Surface Rolling on the Fatigue Strength of Steel," Journal of Applied Mechanics,
Transactions, A.S.M.E., Vol. 57, December, 1935, pp. A-128–136.

Love, R. J., "Cold Rolled Fillets," Engineering, August 8, 1952.

Ford Motor Company; Manufacturing Research Office.

General Motors Corporation (Patent #2,357,515).

Industrial Metal Products Corporation.

International Harvester Company (Patent #2,841,861).

Madison Industries, Inc.

"Shot Peening and Other Surface Working Processes," Supplement to Metals Handbook, Metal
Progress, (ASM), July 15, 1954, pp. 104–108.

Almen, J. O., Mattson, R. O., and Fonda, H. E., Report on "Surface Rolling Treatment."

The Foote-Burt Company, Schraner Division.

Timken Roller Bearing Company, "Cold Rolling of Axle Fillets."

Almen, J. O., "Fatigue Durability of Prestressed Screw Threads," Product Engineering, April, 1951.

Cogsdill Tool Products, Inc.

Stewart, W. C., and Ellinghausen, H. C., "Examination and Test of Failed Port Tail Shaft, USS Norfolk
(DL-1)," U.S. Naval Engr. Exp. Station, R&D Report, 040007AZ(2) NSM-000-003, December 5,
1955.

Dugdale, D. S., "Stress-Strain Cycles of Large Amplitude," J. Mech. Phys. Solids, 1959, Vol. 7, pp. 135–
142. Pergaman Press, London, England.
SAE J811 Revised AUG81

Fuchs, H. O., "Shot-Peening Effects and Specifications," ASTM Spec. Tech. Pub. No. 196, pp. 22–32.
Philadelphia, 1958.

Phillips, A., "Improvement of Fatigue Life of Aircraft Components by Coining," published by ASME, 1961.
Paper 61-AV-35.

Almen, J. O., "Fatigue Loss and Gain by Electroplating," Product Engineering, Vol. 22, No. 6, June, 1951.

Grossman, Nicholas, "Effect of Shot Peening on the Brittle Transition Temperature," Metal Progress, p.
352. September, 1950.

Water, K. T., "Production Methods for Cold Working Joints Subjected to Fretting for Improvement of
Fatigue Strength," Preprint Paper No. 75, ASTM, San Francisco Meeting, 1959.

SAE, Shot Peening Manual, SP-84, New York, 1952.

Timoshenko, S., and Goodier, J. N., Theory of Elasticity, 2nd Ed. New York, McGraw-Hill, 1951.

Timoshenko, S., Strength of Materials, 2nd Ed., Part II: Advanced Theory and Problems. New York:
McGraw-Hill 1941.

Mattson, R. L., and Roberts, J. G., "Effect of Residual Stresses Induced by Strain Peening Upon Fatigue
Strength." Internal Stresses and Fatigue in Metals. Amsterdam, Elsevier, 1958.

Fuchs, H. O., and Mattson, R. L., "Measurement of Residual Stresses in Torsion Bar Springs,"
Proceedings, SESA, Vol. 4, No. 1, 1946.

Almen, J. O., "Fatigue Failures are Tensile Failures," Product Engineering. McGraw-Hill Publishing Co.,
March, 1951.

Machine Design, Penton Publishing Co., October 20, 1958.

Harper, W. A., "Explosive Hardening of Steel Proves Out in Field Service," Iron Age, Vol. 185, No. 4, Feb.
4, 1960, p. 85.

Dermott, R. G., "Progress in Explosive Forming", Metal Progress. November 1959.

Courtesy of L. G. Johnson. Research Laboratories, General Motors Corp. Unpublished.

Thum, A., and Bruder, E., "Shape and Fatigue Strength of Rod Eyes and Similar Constructional Parts,"
Deutsche Kraftfahrtforschung, No. 20, 1939, pp. 1–10.

Design News, Dec. 8, 1958.

Wise, S., "Work-Hardening Bolt Holes in Rail Ends," The Railway Gazette, April 29, 1960, pp. 511–512.

"One Ball, No Errors," Machine Design. Penton Publishing Co., February, 1958.

Almen, J. O., "Brittle Structural Failures with Emphasis on Welded Ships," Product Engineering. April
1953.

Burnheim, H. "Surface Pressing of Notched Test Specimens Consisting of Cr-Mn-V Steel VCV 100,"
Luftfahrtforschung, Vol. 20, No. 1, January 20, 1943.
SAE J811 Revised AUG81

Coombs, A. G. H., Sherratt, F., and Pope, J. A., "An Analysis of the Effects of Shot-Peening Upon the
Fatigue Strength of hardened and Tempered Spring Steel." Proceedings of the International
Conference on Fatigue of Metals, Institution of Mechanical Engineers, London; American Institute of
Mechanical Engineers, New York, 1956. London, William Clowes and Sons, 1956.

Courtesy H. R. Neifert, Timken Roller Bearing Co.


Courtesy G. F. Butz, Aluminum Co. of America.
Courtesy C. W. Cable, Boeing Aircraft Co.

Rosenthal, D., Sines, G., and Zizicas, G., "The Effect of Residual Compression on Fatigue," Welding J.,
Research Suppl., Vol. 28, 1949.

Rosenthal, D., and Sines, G., "Effect of Residual Stress on the Fatigue Strength of Notched Specimens,"
Proc. ASTM, Vol. 51, 1951.

Norton, J. T., Rosenthal, D., and Maloof, S. B., "X-Ray Diffraction Study of the Effect of Residual
Compression on Fatigue of Notched Specimens." Welding J., Research Suppl. 1946.

Surface Stressing of Metals, American Society for Metals, Cleveland, 1947; H. F. Moore, "The Problem
Defined;" W. M. Murray, "Measurement of Surface Stresses;" J. O. Almen, "Fatigue of Metals as
Influenced by Design and Internal Stresses;" O. J. Horger, "Stressing Axles and Other Railroad
Equipment by Cold Rolling;" P. R. Kosting, "Progressive Stress-Damage."

Almen, J. O., "Fatigue Weakness of Surfaces," Product Engineering. McGraw-Hill Publishing Co.,
November 1950.

Almen, J. O., "Torsional Fatigue Failures," Part I, Product Engineering, McGraw-Hill Publishing Co.
September 1951.

Almen, J. O., "Torsional Fatigue Failures," Part II, Product Engineering, McGraw-Hill Publishing Co.
March 1952.

Almen, J. O., "Residual Compressive Stress Strengthens Brittle Materials," Product Engineering.
McGraw-Hill Publishing Co., July 1953.

Mattson, R. L., and Almen, J. O., "Effect of Shot Blasting on the Mechanical Properties of Steel," (NA-
115), Final Report OSRD, 3274, 4825, 6647. Washington, 1945.

Green, W. B., "How Processing Affects Bolt Fatigue Strength," Machine Design. Penton Publishing Co.,
December 1947.

Weibull, W., "The Effect of Decarburization and Other Factors on the Fatigue Strength of Roll-Threaded
Aircraft Bolts," SAAB TN 4, Svenska Aeroplan Aktiebolaget. Linkoping, Sweden, July 1952.

Buckwalter, T. V., and Horger, O. J., "Investigation of Fatigue Strength of Axles, Press-Fits, Surface
Rolling, and Effect of Size," Transactions of the American Society for Metals, Vol. 25, March 1937, p.
229.

Horger, O. J., and Maulbetsch, J. L., "Increasing the Fatigue Strength of Press-Fitted Axle Assemblies by
Surface Rolling," Journal of Applied Mechanics, September 1936, pp. A-91 to A-98.

Horger, O. J., Buckwalter, T. V., and Neifert, H. R., "Fatigue Strength of 5-1/4 in. Diameter Shafts as
Related to Design of Large Parts," Journal of Applied Mechanics, September 1945, pp. A-149 to A-
155.
SAE J811 Revised AUG81

Horger, O. J., and Cantley, W. I., "Design of Crankpins for Locomotives," Transactions ASME, Vol. 68,
1946, pp. A-17 to A-33.

Horger. O. J., "Residual Stress." Handbook of Experimental Stress Analysis, John Wiley & Sons, Inc.,
New York, 1950.

Horger, O. J., "Influence of Fretting Corrosion on the Fatigue Strength of Fitted Members," Symposium
and Fretting Corrosion, Special Technical Publication No. 144, ATM, 1953, pp. 40–51.

Horger, O. J., and Lipson, C., "Automotive Rear Axles and Means of Improving Their Fatigue
Resistance," ASTM Spec. Tech. Publ. 72, "Symposium on Testing of Parts and Assemblies,"
Philadelphia, 1947.

Horger, O. J., "Stresses Imposed by Processing," SAE Quart. Trans. Vol. 5, No. 3, pp. 393–403, July
1951.

Horger, O. J., and Neifert, H. R., "Effect of Surface Conditions on Fatigue Properties," Surface Treatment
of Metals, ASTM, 1941.

Horger, O. J., and Neifert, H. R., "Fretting Corrosion of Large Shafts as Influenced by Surface
Treatments," Symposium on Large Fatigue Testing Machines and Their Results, Special Technical
Publication No. 216, ASTM, 1957, pp. 81–95.

Horger, O. J., and Neifert, H. R., "Correlation of Residual Stresses with Fatigue Strength of Machine
Elements and Related Phenomena," Residual Stresses in Metals and Metal Construction. Reinhold
Publishing Corp., New York, 1954, pp. 219–253.

Neifert, H. R., and Robinson, J. H., "Further Results from the Society's Investigation of Tailshaft
Failures," Transactions, Society of Naval Architects and Marine Engineers, Vol. 63, 1955, pp. 495–
550.

Fuchs, H. O., "Trapped Stresses," Machine Design. Penton Publishing Co., July 1948.

Fuchs, H. O., "Techniques of Surface Stressing to Avoid Fatigue," Metal Fatigue, edited by G. Sines and
J. L. Waisman, University of California Engineering Extension Series. McGraw-Hill Book Co., Inc.,
1959.

Dolan, T. J., "Basic Concepts of Fatigue Damage in Metals," op. cit.

Peterson, R. E., "Fatigue Cracks and Fracture Surfaces—Mechanics of Development and Visual
Appearance," op. cit.

Shook, L. L., Jr., and Long, C. L., "Surface Cold Rolling of Marine Propeller Shafting." paper presented
December 6, 1957, to the Hampton Roads Section, The Society of Naval Architects and Marine
Engineers.

Egger, Walter, and Diamond, Gerald X., "Fillet Rolling," Machine Design. Penton Publishing Co.,
January 5, 1961.

Garwood, M. F., ZurBurg, H. H., and Erickson, M. A., "Correlation of Laboratory Tests and Service
Performance," Interpretation of Tests and Correlation with Service, ASM, 1950.
SAE J811 Revised AUG81

Love, R. J., and Waistall, D. N., "The Improvement in the Bending Fatigue Strength of Production
Crankshafts by Cold Rolling," Report No. 1954/2, published by the Motor Industry Research
Association, Lindley, Warwickshire, England. 1954.

Love, R. J., "Fatigue in Automobiles," Proceedings of the International Conference on Fatigue of Metals,
Institution of Mechanical Engineers, London; ASME, New York; 1956.

Siebel, E., and Gaier, M., "The Influence of Surface Roughness on the Fatigue Strength of Steels and
Non-Ferrous Alloys," translated in The Engineers' Digest. March 1957.

Bibliography on Residual Stress (SP-125), and Supplement I (SP-167), SAE.

Evaluation of Methods for Measurement of Residual Stress, (HS 147), SAE.

Horger, O. J., "Cold Working," ASME Handbook Metals Engineering-Design, 2nd edition, O. J. Horger,
ed. McGraw-Hill, 1964, pp. 264–267.

Metals Handbook, Vol. 3, 8th edition, "Roller Burnishing," pp. 105–107, "Thread Rolling," pp. 130–145.

Developed by the SAE Iron and Steel Technical Committee

You might also like