You are on page 1of 8

Engineering Failure Analysis 34 (2013) 511–518

Contents lists available at ScienceDirect

Engineering Failure Analysis


journal homepage: www.elsevier.com/locate/engfailanal

Short communication

Analysis of a failed spur gear from a Vibro-Hammer


J. Jonck ⇑, G.A. Slabbert
Metals Technology Centre, Advanced Materials Division, Mintek, Private Bag X3015, Randburg 2125, South Africa

a r t i c l e i n f o a b s t r a c t

Article history: This paper presents the findings from the failure analysis of an induction-hardened spur
Received 25 March 2013 gear from a Vibro-Hammer, which failed after two months’ operation. The analysis of
Accepted 22 July 2013 the gear showed that the failure occurred due to tooth-bending fatigue. The premature fati-
Available online 31 July 2013
gue was attributed to incorrect case-hardening practice which was aggravated by poor
machining, misalignment of the gear and incorrect heat-treatment prior to induction-hard-
Keywords: ening. It was recommended to improve the case-hardening, machining, installation and
Failure analysis
heat-treatment procedures in order to increase the service-life of the gear.
Spur gear
Fatigue
Ó 2013 Elsevier Ltd. All rights reserved.
Case-hardening

1. Introduction and background

This paper details the analysis of a failed 295 mm diameter spur gear from a Vibro-Hammer used to install sheet-piles.
The gear was reportedly in service for 2 months, which equates to 700–800 h of service. The grade of material used for the
manufacturing of the gear was not supplied. The rotational speed of the gear was 2100 rpm. Possible operator misuse was
reported, however it was also reported that the lubrication and maintenance of the gear was in strict accordance with man-
ufacturer’s norms. There had been no prior failures or complaints regarding the gears.

2. Material and methods

A detailed visual inspection was undertaken to assess the general quality of the gear and identify all relevant fracture fea-
tures. Magnetic particle inspection (MPI) was performed over the gear profile in order to identify the presence of additional
cracks or defects that could have caused or contributed to failure. MPI was the preferred evaluation technique as dye pene-
trant inspection could not eliminate the ‘‘noise’’ from the coarse machining marks. A sample containing a crack was sec-
tioned from tooth number 20 and submerged in liquid nitrogen before opening the crack under impact loading for
further evaluation. The MPI was performed using Chemserve Systems Adrox 800-3 black magnetic ink in conjunction with
Adrox 8901W white background lacquer.
The chemical composition of the gear was determined using a SPECTRO MAXx LMM04 spark emission spectrometer. The
specimens used for chemical analysis were ground with silicon carbide grinding paper to a 120 grit surface finish in prep-
aration for testing.
Cross-sectional specimens for metallography were sectioned from the gear at tooth number 22 (containing a crack), the
coarse machined area of an intact tooth, case hardened area of an intact tooth and away from the case hardened region. The
specimens were hot-mounted for good edge retention during examination. The specimens were ground with silicon carbide
grinding paper to a 1200 grit surface finish, followed by three successive polishing steps using 6 lm, 3 lm and 1 lm dia-

⇑ Corresponding author. Tel.: +27 11 709 4164, mobile: +27 83 414 3358.
E-mail address: jacoj@mintek.co.za (J. Jonck).

1350-6307/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.engfailanal.2013.07.026
512 J. Jonck, G.A. Slabbert / Engineering Failure Analysis 34 (2013) 511–518

mond suspensions, in order to achieve a sufficient surface finish for metallography. The grinding and polishing were per-
formed using a Struers Abrapol automatic grinding and polishing machine. The specimens were analysed in the as-polished
and etched conditions. The specimens were etched with a 3% Nital solution. The specimens were examined with an Olympus
BX61 optical microscope.
In order to determine the bulk hardness of the gear, Brinell hardness measurements were performed with a Wolpert
Group 3000BLD machine while applying a load of 750 kg using a 5 mm tungsten carbide ball. Micro-Vickers hardness testing
was performed with a Struers Duramin micro-Vickers hardness testing machine in order to determine the depth of the case
at various locations along the tooth profile.

3. Results and discussion

3.1. Visual examination

The gear initially had 36 teeth around the circumference. Tooth number 1 and 8 fractured during service, and were not
received with the remainder of the gear for analysis, as shown in Fig. 1. The heat tint from the case hardening process was
still visible on the surface of the gear, indicating that the gear was induction hardened using the contour hardening (tooth-
to-tooth) method [1], as shown in Fig. 2. The contour-hardening method is preferred over spin-hardening for a gear of this
size and application [1]. The profile of the tint is unusual as it is not uniform around the fillet radii between the teeth. The
tooth roots, including sections of the fillet radii, showed coarse machining marks originating from the hobbing process, as
shown in Fig. 3.
The gear showed no indications of excessive wear. However, the contact patterns were displaced to one side of the gear
teeth, which suggests misalignment. There were no signs of service abuse or lubrication irregularities [1,2].
The received gear was relatively clean and the features on the fracture surfaces were clearly visible. The fracture surface
of tooth number 1 was flat and showed beach marks indicating that the tooth fractured by tooth bending fatigue, as shown in
Fig. 4. The shape of the beach marks and the absence of ratchet marks indicate that the fracture originated on the active flank
from a single initiation point at the surface located away from the centre (co-incident with the side showing wear) within the
coarse machined area of the fillet, as shown in Fig. 4. This is typical of tooth bending fatigue as the misalignment would shift
the peak loading area away from the centre of the gear [2].
The fracture surface of tooth number 8 also showed beach marks originating from the loaded side, similar to tooth num-
ber 1, but differed in that there were two clear origins that formed from multiple initiation sites as indicated by the ratchet
marks, as shown in Figs. 5 and 6. Ratchet marks are ‘‘boundary lines’’ formed between two cracks that propagate at different
planes. This generally either indicates that the nominal stress on the gear is high, or that the stress concentration factor is
high [2]. Both origins were located in the coarse machined region of the fillet radius. However, the fracture surfaces of both
teeth showed small final fracture regions which indicate that the apparent magnitude of stress applied at failure was low [2].
Excessively high stresses would have led to a shear final fracture during tooth bending fatigue even though the applied stress
on the cracked tooth is progressively displaced to adjacent teeth due to the increased deflection of the cracked tooth during
loading [2,3]. This suggests that there may be a stress raiser associated with the failure of tooth number 8.

3.2. Magnetic particle inspection (MPI)

Three cracks were identified on tooth numbers 2, 20 and 22, which were approximately 20 mm, 15 mm and 15 mm in
length, respectively, as shown in Fig. 7. All the cracks were located at the fillet radii on the active flank of the teeth. The cracks
were all singular hairline cracks, indicating that they are probably fatigue cracks. The centre of all the cracks was located

Fig. 1. Photograph of the gear in the as-received condition showing the two fractured teeth.
J. Jonck, G.A. Slabbert / Engineering Failure Analysis 34 (2013) 511–518 513

Fig. 2. Photograph of the gear teeth showing the heat tint from the induction hardening process.

Fig. 3. Photograph of the tooth roots showing the coarse machined regions.

Fig. 4. Photograph of the fracture surface of tooth number 1 showing a flat surface with beach marks originating from a single initiation at the surface in the
coarse machined region.

within the coarse machined region. The cracks at teeth number 20 and 22 were offset to the peak-loaded side of the gear,
while tooth number 2 was not.
The opened crack at tooth number 20 had a thumbnail shape and showed beach marks typical of fatigue, as shown in
Fig. 8. The crack propagated to a depth of 3.5 mm. The surface showed no ratchet marks which indicates the crack origi-
nated from a single initiation point. The initiation site was found to correspond to one of the roots of the folds/laps in the
coarse machined region. No cracks were found adjacent to tooth 8, which is unusual as tooth bending fatigue normally leads
to cracking and/or failure of a number of adjacent gear teeth [2,3].
514 J. Jonck, G.A. Slabbert / Engineering Failure Analysis 34 (2013) 511–518

Fig. 5. Photograph of the fracture surface of tooth 8 showing a flat surface with beach marks originating from two origins at the surface in the coarse
machined region.

Fig. 6. Photograph of the fracture surface of tooth 8 showing beach marks originating from one of the origins at the surface in the coarse machined region
and two ratchet marks.

Fig. 7. Photograph showing the black magnetic particles indicating the presence of a crack on the active flank of tooth number 20.

3.3. Chemical analysis

The results showed that the gear conformed to the BS970 708M40 CrMo steel specification. BS970 708M40 is a through
hardening material grade that is also readily induction hardened, and is often specified for gears applied to this type of ser-
vice environment [2,4,5] (see Table 1).

3.4. Metallography

The polished specimen from tooth number 22 showed that the crack (identified by MPI) was a singular hairline crack of
3.6 mm in depth which is typical for fatigue cracks. The etched bulk microstructure was uniform across the gear and
J. Jonck, G.A. Slabbert / Engineering Failure Analysis 34 (2013) 511–518 515

Fig. 8. Photograph showing the opened crack found on tooth number 20.

Table 1
Results from the chemical analysis of the gear.

Weight% C Si Mn P S Cr Mo
Gear 0.42 0.33 0.74 0.016 0.007 0.99 0.15
BS970 708M40 [4] 0.36–0.44 0.10–0.35 0.70–1.00 <0.035 <0.040 0.9–1.2 0.15–0.25

consisted predominantly of bainite (predominantly upper bainite) and some martensite, which was present at the interden-
dritic regions of primary solidification, as shown in Figs. 9–11. The variation in the microstructure indicates chemical seg-
regation in the form of solute enrichment at the interdendritic regions. A quench and temper pre-heat treatment is
recommended for this steel to produce a tempered martensite microstructure prior to induction hardening [1,6]. The upper
bainitic microstructure could typically lead to a reduced fatigue life due to unfavourable carbide morphology (for a hardness
level below 373 HB) compared to a tempered martensite microstructure [7].
The case microstructure consisted of tempered martensite. However, the case was not uniform across the tooth profile
and decreased in depth from the tooth tip towards the root and no case was present at the root on the transverse section
where the crack initiated, as shown in Figs. 12 and 13. The longitudinal section across the root showed no case hardening
over most of the length, except for a 5 mm region at the centre of the root, which was up to 100 lm in depth. The absence
of a case hardened region at the fillet radius will significantly reduce the fatigue crack resistance, not only due to the lower
tensile strength in this region, but also due to the absence of compressive residual stresses typically induced by case hard-
ening [1,3,8]. The gear also showed a rapid transition from the case to the base microstructure (typically less than 130 lm)
which is typical for case induction hardened gears [8].

3.5. Hardness testing

The average hardness of the gear (256 HB) conformed to the hardness required for BS970 708M40 steel, tempered to both
conditions S and T, as shown in Table 2. The bulk hardness of the gear is below the lower limit of the typical range recom-

Fig. 9. Micrograph showing the bainite (light phase) and martensite (dark phase) general microstructure of the gear.
516 J. Jonck, G.A. Slabbert / Engineering Failure Analysis 34 (2013) 511–518

Fig. 10. Micrograph showing the bainite portion of the microstructure at higher magnification, see Fig. 12.

Fig. 11. Micrograph showing the mixed martensite–bainite portion of the microstructure at higher magnification, see Fig. 12.

Fig. 12. Photograph of the etched metallography specimen showing a reduction in case depth from the tooth tip to the root. Note that the case is completely
absent in the fillet region where the crack initiated.

mended (285–370 HB) for optimum performance and ease of production of induction hardened gears [1]. The ultimate
tensile strength (UTS) was also estimated from the average hardness.
The average case hardness was 653 Hv1 and is higher than the typically recommended value of 500 Hv1 specified for
induction hardening for optimum fatigue resistance [8]. This could lead to notch sensitivity and reduced fatigue life [8]. This
indicates that the tempering temperature applied after hardening could be too low which could lead to a brittle microstruc-
ture [5].
J. Jonck, G.A. Slabbert / Engineering Failure Analysis 34 (2013) 511–518 517

Fig. 13. Micrograph of the crack origin at the fillet radius showing an absence of case hardening.

Table 2
Hardness testing results and estimated tensile strength values.

Hardness (HB) Minimum Maximum Average UTS (MPa) Hardness (HV)


Gear bulk 252 260 256 870a 269b
Gear case – – 621c 2111a 653
Gear fillet – – 262c 891a 276
BS970 708M40 condition S [5] 223 277 – >761 –
BS970 708M40 condition T [5] 248 302 – >837 –
a
Estimated from measured hardness value.
b
Estimated from Brinell hardness number.
c
Estimated from Vickers hardness number.

Fig. 14. Hardness profiles measured from the surface at various locations along the working surface.

The tooth tip showed a significantly lower hardness within the case, compared to the rest of the tooth. The reason for the
lower hardness is probably due to the overlapping of the individual case hardening profiles produced by the inductor which
effectively tempers the martensite from the previous tooth at a much higher temperature than the remainder of the profile.
The case depth (when hardness drops by more than 170 HB or 10 HRC [8]) was found to decrease with increasing distance
from the thread tip up to the start of the fillet radius where no case hardening was found, as shown in Fig. 14. This causes a
significant decrease in strength at the fillet radius which is shown by the estimated UTS values in Table 2.

4. Conclusions and recommendations

The fillet radius on the active flank is a critical area that experiences maximum tensile bending stresses and stress gra-
dients and requires high strength and toughness to resist fatigue crack initiation during operation [1–3]. The gear failed by
tooth bending fatigue which was primarily induced by a lack of case hardening at the fillet radii on the active flank of the
gear leading to insufficient strength. The failure was accelerated by a number of contributing factors including:
518 J. Jonck, G.A. Slabbert / Engineering Failure Analysis 34 (2013) 511–518

 The coarse machine marks at the fillet which introduced additional micro-stress raisers [2,3].
 Misalignment which causes non-uniform loading across the tooth which increases peak stresses experienced by the teeth.
Misalignment is a common contributor and cause of tooth bending fatigue failures [3]. The cracking and fracture which is
predominant at one end of the gear (but not exclusive) also confirms misalignment as a contributor and not a primary
cause.
 Incorrect heat treatment prior to induction hardening. Upper bainite typically has a far inferior resistance to fatigue crack
initiation compared to the recommended tempered martensite microstructure. This is due to the micro-stress concentra-
tions associated with the tips of the alternating upper bainite fingers [7].

Overloading can also contribute to tooth bending fatigue failure. However, the contact pattern shows no irregularities
that would suggest that the gear has been consistently overloaded (i.e. misuse) [3].
The following improvements are recommended in order to increase the service life of the gear:

 The case hardening process should be improved to produce sufficient hardening at the root fillet and a uniform case thick-
ness in order to prevent similar failures in future. The case hardness could also be reduced for optimum fatigue resistance
as discussed above.
 The pre-hardening heat treatment must be altered to produce a tempered martensite microstructure which is optimum
for induction hardening.
 The regions showing the coarse hobbing marks should be machined to produce a surface finish with lower roughness
before case hardening.
 The gears should be tested after assembly to ensure proper alignment.

Acknowledgement

This paper is published by permission of Mintek. The assistance of our colleagues in the Advanced Materials Division
(AMD) at Mintek is gratefully acknowledged.

References

[1] Otto FJ, Herring DH. Gear heat treatment. In: Davis JR, editor. Gear materials, properties, and manufacture. ASM Int.; 2005. p. 16–7.
[2] Alban LE. Failures of gears. In: ASM handbook, vol. 11. ASM International; 1998. p. 586–601.
[3] Fernandes PJL. Tooth bending fatigue failures in gears. J Eng Fail Anal 1996;3:219–25.
[4] Alban LE. Systematic analysis of gear failures. In: Davis JR, editors. Gear materials, properties, and manufacture. ASM Int.; 2005. p. 12–3.
[5] Iron and steel specifications. 6th ed. British Steel Corporation; 1986. p. 187, 189.
[6] Davis JR. Gear materials, properties, and manufacture. ASM Int.; 2005. p. 249–55.
[7] Breen DB, Wene EM. Fatigue in machines and structures – ground vehicles. In: Boyer HE, editor. Atlas of fatigue curves. American Society for Metals;
1986. p. 31.
[8] Stickels CA. Carburizing, friction, lubrication, and wear technology. In: Davis JR, editor. Gear materials, properties, and manufacture. ASM Int.; 2005. p.
62–3.

You might also like