You are on page 1of 16

Engineering Failure Analysis 14 (2007) 85–100

www.elsevier.com/locate/engfailanal

Failure of an intermediate gearbox of a helicopter


S.K. Bhaumik *, M. Sujata, M. Suresh Kumar, M.A. Venkataswamy,
M.A. Parameswara
Failure Analysis Group, Materials Science Division, National Aerospace Laboratories, P.O. Box 1779, Bangalore 560 017, India

Received 13 December 2005; accepted 25 December 2005


Available online 6 March 2006

Abstract

An intermediate gearbox of a helicopter failed resulting in an accident. A systematic failure analysis was conducted to
find out the cause of failure. Examination revealed that fatigue fracturing of the driving gear was responsible for the gear-
box failure. The teeth of the gear were severely damaged by spalling. Fractographic study revealed multiple fatigue crack
initiation at the tooth root regions. It was established that the failure was caused due to improper assembly of the gear. A
detailed analysis of the failure is presented in this paper.
Ó 2006 Elsevier Ltd. All rights reserved.

Keywords: Gear failure; Fatigue; Spalling; Improper assembly; Misalignment

1. Introduction

Gears are essential machine elements designed to transmit motion and power from one mechanical unit to
another. There hardly exists any engineering machine that operates without gears. Various types of gear have
been developed to perform different functions. The major types are spur gears, helical gears, straight and spiral
bevel gears, and hypoid gears. The gear type and the specific design features determine the operating charac-
teristics of a gear.
Failure of gears can occur because of various reasons. Accurate determination of cause of a gear failure
requires identification of failure modes unambiguously. There are various types of gear failures and they
are classified in four major groups: wear, surface fatigue, plastic flow and breakage. Each of these general clas-
ses of failure is subdivided for more accurate and specific identification. A detailed description of the mech-
anisms of gear failures can be found in the literature [1,2]. It is reported [2,3] that fatigue, impact, pitting,
spalling, crushing, scoring, and scuffing account for more than 75% of the gear failure; tooth bending fatigue
and surface contact fatigue being the two most common modes of failure.
Aircraft gearboxes are generally robust and reliable. Most of the failures in these gearboxes occur due to
application errors, of which, misalignment is probably the most common single cause of failure [4].

*
Corresponding author. Tel.: +91 080 2508 6277; fax: +91 080 2527 0098.
E-mail address: subir@css.nal.res.in (S.K. Bhaumik).

1350-6307/$ - see front matter Ó 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engfailanal.2005.12.006
86 S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100

Misalignment can lead to varieties of problems and makes the gears amenable for failure by various mecha-
nisms. However, each type of failure leaves characteristic features on the gear teeth and a detailed examination
of the failed gear provides valuable information to establish the cause of failure.
This paper analyzes failure of an intermediate gearbox of a helicopter. The failure resulted in an accident to
the helicopter. It was established that the failure was caused due to improper assembly.

2. Background

A loud noise was heard when a helicopter was hovering at a height of about 5 m from the ground. Fol-
lowed by this, the helicopter lost control and took a vicious yaw. The crew immediately lowered the collec-
tive in an attempt to bring down the helicopter but failed. The helicopter impacted on to the ground with a
bang resulting in huge damage to the structure. Preliminary investigation revealed that loss of power trans-
mission to the tail rotor was responsible for the accident. On further examination, the tail rotor was found
freely rotating and the problem was traced to the failure in the intermediate gearbox (IGB). The purpose of
the IGB is to transmit rotary motion at an angle corresponding to the angle of the tail boom. It consists of
a pair of bevel gears. The gears have 28 spiral case-hardened teeth each. Each gear rotates on three bear-
ings; two of them are cylindrical roller bearings taking up radial loads, whereas the third bearing is radial
thrust bearing taking up axial loads. The gears were in service for 4463 h since new and the failure took
place 150 h after the last overhaul.

3. Strip examination of the gearbox

The damaged gears and the casing of the IGB are shown in Figs. 1–3. The driving gear had fractured into
three pieces; one part was with the shaft while the other two parts were retrieved from the gearbox casing
(Fig. 1). The driven gear was in one piece but all the teeth were severely damaged (Fig. 2). The damages were
mainly in the form of teeth chewing and deformation, typical of impact damages. There were no cracks in this
gear. The casing of the gearbox was in tact (Fig. 3) but for some metal gouging marks on the inner surface
(Fig. 4). These had occurred due to impact of the broken parts of the driving gear. The damages on the driven
gear and the casing were caused due to failure of the driving gear and they were secondary in nature. The bear-
ings on both driving and driven gear shafts were found freely rotating.

Fig. 1. Failed driving gear; A, B and C denote three fractured parts.


S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100 87

Fig. 2. Damaged driven gear showing teeth chewing.

Fig. 3. Casing of the intermediate gearbox.

4. Driving gear failure

4.1. Visual and stereo-binocular examination

The broken pieces of the driving gear were cleaned and examined for identification of the mode of fracture.
Out of the three pieces, two pieces (marked A and B in Fig. 1) had fallen off in the gearbox casing after frac-
ture, while the third part (C in Fig. 1) was with the gear shaft. These three fractured pieces are shown sepa-
rately in Figs. 5–7. It can be seen that the teeth in the fractured part A were in tact, while the same on the other
two parts (B and C) were completely chewed off and/or fractured from the tooth root region. This indicates
88 S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100

Fig. 4. Inner surface of the gearbox casing showing gouge marks (shinny).

Fig. 5. Photograph showing part A of the fractured driving gear.

Fig. 6. Photograph showing part B of the fractured driving gear.


S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100 89

Fig. 7. Photograph showing part C of the driving gear with the shaft.

that part A of the gear had fractured first and fallen off. The discontinuity in gear had then resulted in impact
loading with the driven gear leading to shaving off of all the teeth of pieces B and C. During this, the second
fracture of the gear had occurred.
The gross fractographic features on all the fracture surfaces were indicative of progressive failure. Beach
marks, typical of fatigue failure were present on the fracture surfaces. Examination revealed fatigue crack ini-
tiation at six different locations. These are summarized in Fig. 8. Tracing back the beach marks, the fatigue
crack origins were identified and these are shown by arrows. In all the cases, the fatigue crack had initiated
at the tooth root region. After initiation, the cracks had propagated in the gear hub progressively over a period
of time before culminating in final rupture. Though there were six fatigue crack origins, the number of major
cracks in the gear was three.
The part A of the gear had fractured first and had fallen off into the gearbox casing. Hence it is presumed
that the damages on the teeth prior to the fracture are preserved on this part. Examination of these teeth
would indicate the state of stress prevailing on the gear before failure occurred.
Examination revealed severe wear and material flow on the loading flank of the teeth. A close-up view is
shown in Fig. 9. This is typical of destructive wear and indicative of excessive load on the gear teeth, in gen-
eral. On the non-loading flank, minor polishing wear was observed on a localized region (Fig. 10). As far as
the functioning of the gear is concerned, it rotates in one direction, i.e., clockwise direction. Therefore, the
wear on the non-loading flank is unusual and can take place only when the gears do not mesh properly. Sim-
ilar polishing wear was also seen on the non-loading flank of the driven gear.

4.2. Scanning electron fractography

The fracture surfaces were examined under a scanning electron microscope (SEM) for detailed study.
Fig. 11 shows one of the fatigue crack origins in part A of the fractured driving gear. Well-defined beach
marks typical of fatigue can be seen. From the orientation of the beach marks, the fatigue crack origin could
be determined unambiguously and it is shown by an arrow. The fatigue crack had initiated at the tooth root at
about 3.5 mm away from the edge. At higher magnifications, closely spaced fatigue striations (Fig. 12) and
dimples (Fig. 13) were seen in the progressive crack propagation region and final rupture region respectively.
There were no metallurgical and/or mechanical abnormalities at any of the fatigue crack origins. The fracto-
graphic features were similar in all the six fatigue crack origin regions.
90 S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100

Fig. 8. Photographs showing six fatigue crack origins (arrows) on the driving gear; I, II and III denote three major cracks.

Fig. 9. Wear pattern on the loading flank.

The loading flank of the teeth on the fractured part A was examined. Craters and subsurface cracks were
observed on the entire flank surface (Fig. 14). The craters were formed due to spalling. Spalling is the culmi-
nation of surface contact fatigue and generally, indicates high stresses on the gear teeth.
S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100 91

Fig. 10. Polishing wear on a localized region of the non-loading flank.

Fig. 11. SEM fractograph showing a fatigue crack origin (arrow) in part A of the fractured driving gear.

4.3. Microstructure, composition and hardness

A specimen from the gear tooth was metallographically prepared and observed under an optical microscope
in etched condition. Fig. 15 shows the microstructure of the core. It consists of tempered martensitic structure.
The microstructure of the case showed a satisfactory carburized structure (Fig. 16). Compositional analysis
was carried out at several locations of the gear and the average value is given in Table 1. The gear material
was found to conform to SAE 9130H steel specification.
The hardness survey on the gear tooth was carried out using a microhardness tester at a load of 500 g. The
hardness distribution across the tooth thickness at the pitch line is shown in Fig. 17. The core and surface
hardness were measured to be 390 and 700 HV, respectively. The case depth at 550 HV was measured to
be about 0.6 mm.
92 S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100

Fig. 12. SEM fractograph showing fatigue striations.

Fig. 13. SEM fractograph showing dimple rupture in the final fracture region.

4.4. Analysis of driving gear failure

Fractographic features showed that the driving gear has failed by fatigue. There were three major
cracks and six fatigue crack origins. All the fatigue cracks had originated at the gear tooth region and
propagated progressively over a period of time before culminating in final rupture. There were no metal-
lurgical and/or mechanical abnormalities at the fatigue crack origin region. Metallographic study and
hardness survey showed that the microstructure, case depth and core hardness of the gear were
satisfactory.
Examination of the loading flank of the driving gear teeth showed excessive wear and deformation (see
Fig. 9). This type of wear usually results when the load on the flank is excessively high and is, generally
S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100 93

Fig. 14. SEM photograph showing craters and subsurface cracks.

Fig. 15. Optical microstructure of the core showing tempered martensitic structure.

referred to as destructive wear. Destructive wear is caused by direct tooth contact, and is not related to abra-
sives or corrosion. Such wear occurs over most of the gear tooth face except at the pitch line [1].
The severe spalling observed on the loading flank of the driving gear (see Fig. 14) results from subsurface
fatigue due to high Hertzian stresses, and is indicative of extremely high operating loads. Spalling is nothing
but the manifestation of surface contact fatigue and generally occurs at the region of maximum surface con-
tact stresses. In this mechanism of failure, initially, minute sub-surface cracks are generated at the region of
maximum shear stress, which is below the contacting surface. With time, these cracks grow in size in a direc-
tion parallel to the surface by fatigue mechanism due to repeated stress cycles. The individual cracks then join
together and exit to the surface resulting in detachment of a layer of material leaving behind a crater on the
flank [3,5–7]. Once the spalling starts, it progresses further at a relatively faster rate.
94 S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100

Fig. 16. Optical microstructure of the case.

Table 1
Chemical composition (wt.%) of driving gear
Element Driving gear SAE 9310H steel
C 0.11 0.07–0.13
Mn 0.38 0.4–0.7
Si 0.27 0.15–0.35
Ni 3.25 2.95–3.55
Cr 1.6 1.0–1.45
Mo – 0.08–0.15
Fe Balance Balance

750
700
650
Hardness (HV)

600
550
500
450
400
350
300
0 0.5 1 1.5 2 2.5
Distance from surface (mm)

Fig. 17. Hardness profile on the cross-section of a typical tooth along the pitch line.

From the above, it is evident that the driving gear had experienced high loads in service. These loads, in
turn, resulted in severe stresses at the tooth root and induced fatigue crack initiation. After initiation, the
cracks have propagated in the hub of the gear, which indicates that the load on the gear, in general, was sub-
stantially high.
S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100 95

It was reported that the gear was in service for about 4463 h since new and the failure took place 147 h after
the last overhaul. During the overhaul, the gear was subjected to magnetic particle test for crack detection and
it was cleared for reassembly. This means that the fatigue cracks had initiated after the last overhaul, i.e., dur-
ing the last 147 h in service. Microstructural examination and hardness survey showed that the gear did not
have any metallurgical abnormalities. This is in agreement with the fact that no failure was reported in the first
4316 h of service.
Examination of the non-loading flank of the driving gear showed polishing wear on a part of the tooth face
(see Fig. 10). This implies improper meshing of the gear teeth, possibly resulting from misalignment. The mis-
alignment would have occurred during the assembly after the last overhaul of the IGB. Any misalignment in
the gear train, including the drive shaft, may result in extraordinary high loads on the gear, and may, there-
fore, lead to rapid failure by fatigue mechanism.

5. Inspection of bearings and shafts

In view of the above findings, the gear assemblies were dis-assembled and subjected to detailed inspection to
look for abnormalities present, if any. Observations revealed damages on the driving shaft and one of the

Fig. 18. Photograph of the roller bearing showing damages on the face of the inner race.

Fig. 19. Close-up view of the damages on the face of the inner race of roller bearing.
96 S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100

roller bearings. It may be noted that this particular roller bearing was located adjacent to the gear where the
damages were noticed on the shaft.

5.1. Roller bearing

Mechanical damages in the form of adhesion and dislodgement of material were observed in patches
at a few places on the face of the inner race of the bearing in contact with the flange of the gear
(Fig. 18). Under the stereo-binocular microscope, these regions appeared rough, typical of worn surface
(Fig. 19).
A sample containing the damages was cut from the inner race and examined under the SEM. A typical
appearance of the damaged regions is shown in Fig. 20. These regions showed loss of material by repeated
microfracturing. Semiquantitative chemical analysis by EDX showed transfer of material from the gear flange

Fig. 20. SEM photograph showing the worn surface on the face of the inner race of the roller bearing.

Fig. 21. EDX spectrums taken on: (a) base material and (b) damaged regions on the roller bearing inner race; note presence of Ni in the
damaged region.
S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100 97

onto the bearing face (Fig. 21). It may be noted that the bearing is made of Fe–Cr steel while the shaft/gear is
made of Fe–Ni–Cr steel.
Similar damages were observed on the bore surface of the inner race (Fig. 22). Microcracks, spalling and
transfer of material from the shaft onto the bearing were observed in these damaged regions as well (Figs. 23
and 24).
It may be noted that except for the damages described above, no other damages were observed in the bear-
ing. The rollers were in good condition and the bearing was found freely rotating.

Fig. 22. Photograph showing the damages on the bore surface of the roller bearing.

Fig. 23. SEM photograph showing a typical damage on the bore surface of the roller bearing.
98 S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100

5.2. Shaft

The damages on the shaft and the gear flange are shown in Figs. 25–27. These damages were similar to
those observed on the inner race of the roller bearing. SEM examination confirmed that these damages are
identical to those seen on the bearing inner race.

5.3. Analysis of bearing and shaft damages

Examination revealed damages in one of the roller bearings and the driving shaft. The nature of damages is
typical of that observed in fretting failure. Fretting is a wear phenomenon that occurs between two mating

Fig. 24. SEM photograph showing microcracks and spalling on the bore surface of the roller bearing.

Fig. 25. Photograph of the driving shaft; A and B denote the damaged regions on the shaft and the gear flange, respectively.
S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100 99

Fig. 26. Close-up view of the damaged region on the shaft (marked A in Fig. 24).

Fig. 27. Close-up view of the damaged region on the flange of the gear (marked B in Fig. 24).

surfaces and is adhesive in nature. Generally, fretting occurs between two tight-fitting surfaces such as the
present one that are subjected to a cyclic, relative motion of extremely small amplitude. It may be noted that
the fretting surfaces are never intended to undergo relative motion [8].
Minute oscillatory motion between the mating surfaces usually results in a series of elliptical-shaped blem-
ishes that develop centrally located cavities as the fretting progresses. This has been observed on the driving
shaft as well as on the bore surface of the roller bearing (see Figs. 23–26). Blemishes that do not encircle the
shaft indicate that the bearing loads are absorbed by a limited-contact area. In the present case, machining
marks were preserved on majority of the bore surface (see Fig. 23) indicating such limited contacts. The
limited contacts between the mating surfaces can occur because of improper tolerances or ovality or excessive
surface undulations on the components.
100 S.K. Bhaumik et al. / Engineering Failure Analysis 14 (2007) 85–100

6. Cause of the gearbox failure

It is established that one of the roller bearings was not assembled properly on the driving shaft. There was
relative movement between the shaft and bore of the bearing resulting in fretting damage. Once sufficient
clearance was established due to fretting, the driving gear probably began pounding (low amplitude) on the
driven gear causing excessive load on the teeth. This in turn would have resulted in excessive wear on the load-
ing flank as well as fatigue crack initiation in the gear. The polishing wear seen on a localized region of the
non-loading flank confirms the improper meshing and/or misalignment of the gears resulting from improper
assembly.

Acknowledgements

The authors acknowledge the contributions of Mr. M. Madan in metallographic sample preparation,
microstructural study and hardness measurements. Thanks are due to Ms. Kalavati for scanning electron
microscopy support. The authors are thankful to Director, NAL for permission to publish this work.

References

[1] Failures of gears. Failure analysis and prevention, ASM handbook, vol. 11. Metals Park (OH): American Society for Metals; 1986. p.
507.
[2] Alban LE. Why gears fail. In: Source book in failure analysis. Metals Park (OH): American Society for Metals; 1974. p. 84.
[3] Fernandes PJL. Tooth bending fatigue failures in gears. Eng Fail Anal 1996;3:219.
[4] Asi O. Fatigue failure of helical gear in a gearbox. Eng Fail Anal 2005, doi:10.1016/j.engfailanal.2005.07.020.
[5] Wulpi DJ. Understanding how components fail. Metals Park (OH): American Society for Metals; 1985. p. 94.
[6] Ding Y, Rieger NF. Spalling formation mechanism for gears. Wear 2003;254:1307.
[7] Fernandes PJL, McDuling C. Surface contact fatigue failure in gears. Eng Fail Anal 1997;4:99.
[8] Failures of rolling-element bearings. Failure analysis and prevention, ASM handbook, vol. 11. Metals Park (OH): American Society
for Metals; 1986. p. 416.

You might also like