Sie sind auf Seite 1von 14

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/228422578

Dynamic modeling and analysis of a spur


planetary gear involving tooth wedging
ARTICLE in EUROPEAN JOURNAL OF MECHANICS - A/SOLIDS APRIL 2010
Impact Factor: 1.68 DOI: 10.1016/j.euromechsol.2010.05.001

CITATIONS

READS

37

384

2 AUTHORS:
Yi Guo

Robert G. Parker

National Renewable Energy Laboratory

Virginia Polytechnic Institute and State Uni

21 PUBLICATIONS 138 CITATIONS

136 PUBLICATIONS 2,617 CITATIONS

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

SEE PROFILE

Available from: Robert G. Parker


Retrieved on: 01 February 2016

European Journal of Mechanics A/Solids 47 (2014) 45e57

Contents lists available at ScienceDirect

European Journal of Mechanics A/Solids


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

Nonlinear dynamics and stability of wind turbine planetary gear sets


under gravity effects
Yi Guo a, *, Jonathan Keller a, Robert G. Parker b
a
b

National Wind Technology Center, National Renewable Energy Laboratory, Mail Stop: 3811, 15013 Denver West Parkway, Golden, CO 80401-3305, USA
Department of Mechanical Engineering, Virgina Tech, USA

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 23 August 2012
Accepted 20 February 2014
Available online 12 March 2014

This paper investigates the dynamics of wind turbine planetary gear sets under the effect of gravity using
a modied harmonic balance method that includes simultaneous excitations. This modied method
along with arc-length continuation and Floquet theory is applied to a lumped-parameter planetary gear
model including gravity, uctuating mesh stiffness, bearing clearance, and nonlinear tooth contact to
obtain the dynamic response of the system. The calculated dynamic responses compare well with time
domain-integrated mathematical models and experimental results. Gravity is a fundamental vibration
source in wind turbine planetary gear sets and plays an important role in the system dynamic response
compared to excitations from tooth meshing alone. Gravity causes nonlinear effects induced by tooth
wedging and bearing-raceway contacts. Tooth wedging, also known as a tight mesh, occurs when a gear
tooth comes into contact on the drive-side and back-side simultaneously and it is a source of planetbearing failures. Clearance in carrier bearings decreases bearing stiffness and signicantly reduces the
lowest resonant frequencies of the translational modes. Gear tooth wedging can be prevented if the
carrier-bearing clearance is less than the tooth backlash.
2014 Elsevier Masson SAS. All rights reserved.

Keywords:
Dynamics
Wind turbine
Planetary gear

1. Introduction
The National Renewable Energy Laboratory (NREL) Gearbox
Reliability Collaborative (GRC) was established by the U.S. Department of Energy in 2006. Its key goal is to understand the root causes
of premature gearbox failures (Musial et al., 2007) through a
combined approach of dynamometer testing, eld testing, and
modeling (Link et al., 2011), resulting in improved wind turbine
gearbox reliability and a reduction in the cost of energy. As a part of
the GRC program, this paper investigates gravity-induced dynamic
behaviors of planetary gear sets in wind turbine drivetrains that
could reduce gearbox life. Planetary gear sets have been used in
wind turbines for decades because of their compact design and
high efciency. Despite these advantages, planetary gear sets
generate considerable noise and vibration. Vibration causing high
dynamic loads may result in gear tooth and bearing failures (Musial
et al., 2007). Fatigue failures are a concern in long life-cycle applications. Analyzing the dynamics of wind turbine planetary gear

* Corresponding author. Tel.: 1 303 384 7187; fax: 1 303 384 6901.
E-mail addresses: yi.guo@nrel.gov, guo.83@buckeyemail.osu.edu (Y. Guo).
http://dx.doi.org/10.1016/j.euromechsol.2014.02.013
0997-7538/ 2014 Elsevier Masson SAS. All rights reserved.

drivetrains is important in improving gearbox life and reducing


noise and vibration.
The majority of wind turbines use a horizontal-axis conguration; thus, gravity becomes a periodic excitation source in the
rotating carrier frame. Prior study of gravity by the authors was
performed with a static analysis and focused on the effect of gravity
upon bearing force and tooth wedging in a planetary spur gear set
(Guo and Parker, 2010). It was found that tooth wedging, an
abnormal contact situation where the tooth is in contact with both
the drive-side and back-side anks simultaneously, was caused by
gravity. Tooth wedging increases planet-bearing forces and disturbs
load sharing among the planets, which could lead to premature
bearing failure.
Signicant in-plane translational gear component motions in
planetary systems lead to tooth wedging. It is the combined effect
of gravity and bearing clearance nonlinearity. Bearing clearance
results in greater translational vibration, while gravity is the
dominant excitation source causing the large motions that lead to
tooth wedging. For heavy planetary gear sets, tooth wedging is
likely to occur. Tooth wedging in planetary gear sets leads to unequal load sharing and excessive planet-bearing loads by disturbing
the symmetry of the planet gears. This may cause bearing failure
and tooth damage (Guo and Parker, 2010).

46

Y. Guo et al. / European Journal of Mechanics A/Solids 47 (2014) 45e57

Research on tooth separation is well-established for automotive


and helicopter applications. Tooth separation was observed in spur
gear pair experiments (Blankenship and Kahraman, 1996). Botman
(1976) experimentally observed tooth separation in planetary gear
sets. Using nite element and lumped-parameter models,
Ambarisha and Parker (2007) predicted tooth separation and other
nonlinear phenomena in a planetary gear set in a helicopter
gearbox. Velex and Flamand (1996) investigated tooth separation at
critical speeds. Bahk and Parker (2011) derived closed-form solutions for the dynamic response of planetary gear sets with tooth
separation based on a purely torsional model.
Nonlinear dynamics induced by bearing clearance has been
studied for relatively small geared systems. Kahraman and Singh
(1991) observed chaos in the dynamic response of a geared rotorbearing system with bearing clearance and backlash. Gurkan and
Ozguven (2007) studied the effects of backlash and bearing clearance in a geared exible rotor and the interactions between these
two nonlinearities. Guo and Parker (2012a) investigated the
nonlinear effects and instability caused by bearing clearance in
helicopter planetary gear sets. Dynamic effects of bearing clearance
in wind turbine planetary gear sets have not been studied in the
past because the wind turbine operating speed was believed to be
well below the frequency range of drivetrain dynamics. However,
bearing clearance reduces some gearbox resonances signicantly.
The nite element program developed by Vijayakar (1991) uses
a combined surface integral and nite element approach to capture
tooth deformation and contact loads in geared systems. This nite
element model includes bearing clearance, tooth separation, tooth
wedging, uctuating mesh stiffness, and gravity. Numerical integration is widely adopted to compute dynamic responses of mechanical systems in the time domain. Ambarisha and Parker (2007)
used numerical integration to study nonlinear dynamics and the
impacts of mesh phasing on vibration reduction of planetary gear
sets. Velex and Flamand used numerical integration results of a
planetary gear set with time-varying mesh stiffness as a benchmark
to evaluate results from a Ritz method.
The harmonic balance method (Thomsen, 2003) calculates
nonlinear, frequency domain, steady-state response of mechanical
systems. Zhu and Parker (2005) used this method to study clutch
engagement loss in a belt-pulley system. Al-shyyab and Kahraman
(2005a,b) investigated primary resonances, subharmonic resonances, and chaos in a multimesh gear train caused by uctuating
gear mesh stiffness. Bahk and Parker (2011) employed harmonic
balance to analyze planetary gear dynamics based on a purely
rotational model. Use of the harmonic balance method reduces
computational time for lightly damped or physically unstable
systems by avoiding the long transient decay time before a steady
state is reached. Compared to numerical integration and nite
element analysis, which are widely adopted approaches to
compute dynamic responses, the computation time of the harmonic balance method is oneetwo orders of magnitude lower.
Harmonic balance often employs arc-length continuation (Nayfeh
and Balachandran, 1995) and Floquet theory (Raghothama and
Narayanan, 1999; Seydel, 1994) to calculate nonlinear resonances in the dynamic response, including unstable solutions that
numerical integration and nite element analysis are unable to
obtain. The established harmonic balance formulation is only
suitable for systems with one fundamental excitation frequency
and its higher harmonics. However, wind turbine drivetrains have
simultaneous internal and external excitations, including uctuating mesh stiffness, gravity, bending-moment-induced excitations in the rotating carrier frame, wind shear, tower shadow, and
other aero-induced excitations.
The major objectives of this study are to: 1) develop a modied
harmonic balance method to obtain the dynamic response of wind

turbine planetary gear sets considering gravity, uctuating mesh


stiffness, bearing clearance, and nonlinear tooth contact; 2) validate
the proposed method by comparing the calculated results against
experimental data and a numerical integration approach; and 3)
investigate the gravity-induced dynamic behaviors using the
developed approach, which includes tooth wedging, tooth contact
loss, and bearing-raceway contacts.
2. Gearbox description
This study investigates both a 750-kW wind turbine planetary
gear (PG-A) used by the GRC (Link et al., 2011) and a 550-kW wind
turbine planetary gear (PG-B) (Guo and Parker, 2010; Larsen et al.,
2003; Rasmussen et al., 2004). These drivetrains have a main
bearing that supports the main shaft and rotor weight, and two
trunnion mounts that support the gearbox. These two gearboxes
have similar congurations and are representative of the majority
of three-point-mounted wind turbine drivetrains.
PG-A is congured in a helical planetary gear arrangement with
two parallel stages as shown in Fig. 1. Within the gearbox, there are
two cylindrical roller bearings supporting the carrier, each with
275 mm of clearance. The planetary gear set is arranged in an inphase bridged carrier design with three equally spaced planets.
The sun pinion shaft is connected to the intermediate stage through
a spline joint that partially oats the sun pinion. The ring gear is
bolted to the front and rear of the gearbox housing. The rated torque is 322,610 Nm and the rated speed of the input shaft is 22.2 rpm
(Guo et al., 2012).
PG-B is congured in a spur planetary gear arrangement with
two parallel stages. Like PG-A, PG-B has three equally spaced
planets. The rated torque is 180,000 Nm and the rated speed is
30 rpm. Additional key parameters of these two gearboxes are
listed in Tables 1 and 2. A large ring gear mass in these designs
is a result of a typical wind turbine gearbox arrangement
whereby much of the gearbox housing is rigidly connected to
the ring.
3. Mathematical models
3.1. Lumped-parameter model for planetary gear sets
A previously developed and validated lumped-parameter model
was adopted for this paper (Guo and Parker, 2010, 2012a). As
depicted in Fig. 2(a), the carrier, ring, sun, and planets are rigid
bodies, each having two translational and one rotational degree of
freedom. The carrier rotating frame is used as the general coordinates for all of the components of planetary gear sets. This twodimensional model has 3(N 3) degrees of freedom, where N is the
number of planets. The model includes gravity, uctuating mesh
stiffness, bearing clearance, tooth contact loss, and tooth wedging.

Fig. 1. The GRC gearbox (PG-A) conguration.

Y. Guo et al. / European Journal of Mechanics A/Solids 47 (2014) 45e57


Table 1
System parameters for the 750-kW Gearbox Reliability Collaborative (GRC) planetary gearbox (PG-A). The ring mass includes the gearbox housing and parallel stages.
Sun
Mass (kg)
Moment of Inertia (kg-m2)
Number of Teeth
Pitch Diameter (mm)
Root Diameter (mm)
Average Mesh Stiffness (N/m)
Bearing Stiffness (N/m)
Carrier-Bearing Stiffness (N/m)
Carrier-Bearing Clearance (mm)
Torsional Support Stiffness (N/m)

Ring

Carrier

Planet

181.6
2633
759.9
104
3.2
144.2
59.1
3.2
21
99
e
39
215.6
1016.4
e
400.4
186.0
1047.7
e
372.9
9
9
ksp 16.9  10 , krp 19.2  10
100
102  106 5  109 6.8  109
3.2  109
0.275
45.8  106 57.4  106 0
0

Bearings are modeled using circumferentially distributed radial


springs with a uniform clearance. This study focuses on the carrier
bearing with clearance as shown in Fig. 2(b), which has dynamic
effects on low-speed resonances (Guo and Parker, 2012a).
The coordinates are shown in Fig. 2(a). Throughout this paper,
the subscripts c, r, s, p denote the carrier, ring, sun, and planet; the
subscripts B, g, m denote the bearing, gravity, and gear mesh; and
superscripts b, d denote drive-side and back-side tooth contact.
Translational displacements in the x and y directions, xw, yw, w c,
r, s, are assigned to the carrier, ring, and sun, respectively, with
regard to the rotating carrier frame. The origin O is at the center of
the planetary gear set. The radial and tangential displacements of
the jth planet are denoted by xj, hj, j 1, ., N with respect to the
carrier and oriented for each planet as shown in Fig. 2(a). The
rotational displacements are uv rvqv, v c, r, s, 1, ., N, where qv is
the rotation in radians and rv is the base circle radius for the sun,
ring, and planets and the radius to the planet center for the carrier.
The masses and moments of inertias of the carrier, ring, sun, and
planets are denoted by mk, Ik, k c, r, s, p. Quantities kwx, kwy, kwu
denote the bearing stiffnesses of the carrier, ring, and sun supports
in x, y, and u directions. The torsional stiffnesses of the carrier, ring,
2 , where k
and sun supports equal kwu rw
wu is the torsional stiffness
with units of force/length and rw, w c, r, s is the base radius. The
jth planet-bearing stiffness is kpj. The mesh stiffnesses at the jth
sun-planet and ring-planet mesh are ksj, krj. The radial stiffnesses of
the bearing that connects the carrier to ring gear in the x and y
directions are kcrx, kcry. The nondimensionalized equations of motion of planetary gear sets are

~ 0 ~f d s; z ~f b s; z ~f s; z F
~ s
~ 00 Cz
Mz
B
m
m
x ~
M ~
C ~d
f dm ~b
f bm
; M ; C

; fm
; fm
;
2
L
M
MLu
MLu
MLu2
~f f B ; F
~ F
B
MLu2
MLu2

(1)

where s ut and z x/L, where u is characteristic frequency, L is


the characteristic length, and M is the characteristic mass. Derivations of the mass matrix M and the displacement vector x are
detailed in Guo and Parker (2010). C U1 T diag2zn Un U1
where zn are the damping ratios and Un are the natural frequencies
of the linear system where all bearings are in contact and the mesh
stiffnesses are averaged over a mesh cycle; U is the orthonormalized modal matrix (UTMU I). The nonlinear forces fall into
three categories: the drive- and back-side tooth mesh force vectors
f dm and f bm , and bearing force vector fB (Guo and Parker, 2010). The
external force F(t) Fs Fg(t). Fs includes the static torques applied

47

to each component. Fg(t) is the gravity force vector. Gravitational


force acting on the carrier, ring, sun, and planets is periodic in the
rotating carrier frame, resulting in a fundamental external excitation source.

h
iT
y
h
x h
x
x
x y
x y
Fg t fcg
; fcg ; 0; frg
; frg ; 0; fsg
; fsg ; 0; f1g
; f1g ; 0; .; fNg
; fNg ; 0

(2)

x ; f w c; r; s and f x ; f j 1; .; N denote the


Quantities fwg
wg
jg jg
gravity force acting on the carrier, ring, sun, and planets 1eN. These
are
y

x m g sinU t
fwg
w
c
y
mw g cosUc t;
fwg

w c; r; s



fjgx mj g sin Uc t jj


h
fjg mj g cos Uc t jj ;

j 1; .; N

(3)

(4)

where the variable Uc um/Nr (if the ring is xed) denotes the
carrier rotation frequency. um denotes the mesh frequency. Nr denotes the number of teeth on the ring.
The mesh stiffness uctuates as the number of teeth in
contact changes and is also an important internal excitation
source of geared systems. These excitations are included through
time-varying mesh stiffnesses. The mesh stiffnesses are calculated using the Calyx program (Nayfeh and Balachandran, 1995).
Its mean amplitudes are listed in Tables 1 and 2 (Appendix). This
program analyzes gear tooth contact and rolling element contact
by using a combined analytical/nite element analysis detailed
in Vijayakar (1991). Its results have been compared against
studies of gear dynamics (Singh, 2010, 2011; Guo and Parker,
2012b).
3.2. Finite element model
The two-dimensional nite element model includes 36 quadrilateral elements per tooth, four nodes per element, and each node
has three degrees of freedom. It uses a combined surface integral
and nite element method to capture tooth deformation and contact loads in geared systems (Kahraman and Blankenship, 1994).
The software developed by Vijayakar (1991) intrinsically evaluates
time-varying tooth contact forces that are specied externally with
conventional simulation tools. Fluctuating mesh stiffness over a
mesh cycle, tooth wedging, and tooth separation are included
intrinsically in the nite element model. The nite element model
also includes clearance nonlinearity at the carrier-ring bearing.
Other bearings are modeled as linear stiffnesses without clearance
in the nite element and analytical models. The nite element
approach has been validated by experiments of geared systems
(Parker et al., 2000a,b; Kahraman and Vijayakar, 2001). It is used to
benchmark the established lumped-parameter model when
experimental data is unavailable.
4. Extended harmonic balance method
The extended harmonic balance method is used to obtain the
dynamic responses of the model in Eq. (1). The formulation includes two excitation sources with excitation frequencies U1 and
U2. The coupling effects between these two excitations are
considered by including their side bands. Other excitation sources
can be considered in a similar way. The response z is expanded
into a Fourier series and assumed to include the R1, R2, R3R4R5, and
R6R7R8 harmonics of excitation frequencies U1 and U2 and their

48

Y. Guo et al. / European Journal of Mechanics A/Solids 47 (2014) 45e57

Table 2
System parameters for the 550-kW planetary gearbox (PG-B). The ring mass includes the gearbox housing and parallel stages.
Sun
Mass (kg)
Moment of Inertia (kg-m2)
Number of Teeth
Pitch Diameter (mm)
Root Diameter (mm)
Average Mesh Stiffness (N/m)
Bearing Stiffness (N/m)
Gearbox Trunnion Stiffness (N/m)
Carrier-Bearing Stiffness (N/m)
Carrier-Bearing Clearance (mm)
Torsional Support Stiffness (N/m)

Ring

Carrier

Planet

51
4000
1330
114
61.1
2484
314.7
51.9
16
68
e
26
224
952
e
364
202
980
e
329
9
9
ksp 3.95  10 , krp 5.29  10
100
e
4  109 5.3  109
126  106
3.6  109
1
0
3  106 24.4  106 0

side bands mU1 lU2 and pU1  qU2. Each component zh in z then
has a total of 2(R1 R2 R3R4R5 R6R7R8) 1 terms and is
expressed as

2
6
G6
4

G1 G2

R4 P
R3
P
m1 l1

G3 m;l;R5

R7 P
R6
P
p1 q1

G4 p;q;R8 7
5
n
(7)

The response derivatives are then transformed into

z00 GU21 A1 z  GU22 A2 z  G


G

R7 X
R6
X

R4
R3
P
P
m1 l1

mU1 lU2 2 A3 z

pU1  qU2 2 A4 z

p1 q1

zh zh;1

R1 
X

zh;2i cos iU1 s zh;2i1 sin iU1 s

z0 GU1 B1 z GU2 B2 z G

i1

i1 m1 l1

zh;2ai;m;l1 sin imU1 lU2 s

R7 X
R6
X

pU1  qU2 B4 z

(8)

p1 q1

i1

R5 X
R4 X
R3 n
X
zh;2ai;m;l cos imU1 lU2 s

mU1 lU2 B3 z

m1 l1

R2 h
i
X

zh;2iL1 cos jU2 s zh;2iL1 1 sin iU2 s

R4 X
R3
X

(5)

where the operators A and B are also dened in Appendix A.


The nonlinear force vectors, detailed in Guo and Parker (2012a),
are transformed as

R8 X
R7 X
R6 n
X
zh;2bi;p;q cos ipU1  qU2 s
i1 p1 q1

o
zh;2bi;p;q1 sin ipU1  qU2 s

~f d Gf d ;
m
m

where a(i, m, l) l L2 (m  1)R3 (i  1)R3R4 and b(i, p,


q) q L3 (p  1)R6 (i  1)R6R7. L1 R1, L2 R1 R2,

z1;1 ; .; z1;2L1 1 ; .; z1;2L2 1 ; .; z1;2L3 1 ; .; z1;L4 ; .;


|{z} |{z} |{z}

uN U2

uN U1 ;U2

L3 R1 R2 R3R4R5, L4 2(R1 R2 R3R4R5 R6R7R8) 1.


The time domain is discretized into n  1 evenly distributed
time intervals [s1, ., sn]. Each component zh in the response z is
extended into a vector in the time domain as zh s1 ; .; zh sn T .
The response vector transforms into z Gz by dening a function
G that maps the response from the frequency domain to the time
domain. Additional terms G1eG4 are dened in Appendix A.

~f Gf ;
B
B

~ GF
F

(9)

By substituting Eqs. (8) and (9) into the equations of motion, Eq.
(1) yields

3T

6
7
6
7
xc U1
xc U2
xc U1 ;U2
7
z 6
6 z3N3;1 ; .; z3N3;2L 1 ; .; z3N3;2L 1 ; .; z3N3;2L 1 ; .; z3N3;L 7
1
2
3
4
4
5
|{z} |{z} |{z}
uN U1

~f b Gf b ;
m
m

(6)

Y. Guo et al. / European Journal of Mechanics A/Solids 47 (2014) 45e57

82
R4 X
R3
<
X
2
G 4  U21 MA1  U22 MA2 
mU1 lU2 MA3
:

and JB are described in Guo and Parker (2012a). The


Jacobian matrix associated with back-side tooth loads, Jbm , is

m1 l1

R7 X
R6
X

pU1  qU2 MA4 U1 B1 U2 B2


2

p1 q1

R4 X
R3
X

mU1 lU2 CB3

(10)

m1 l1

R7 X
R6
X

fB

d
fm

b
fm


F

K  U1 MA1  U2 MA2 
2

9
=

Lm;b
r1;i

Lm;b
r2;1

R4 X
R3
X

mU1 lU2 CB3

m1 l1

mU1 lU2 MA3


2

2 P Lb

c1;i

Lc;B

6
6
6
6
6
vFB
G6
H
6
6
vz
6
6
6
4 Symmetric

Lc;B
Lbr Lc;B
Lbs

Lbc2;1 . Lbc2;N 3
7
7
.
7
7
7
.
7
7
7
0 7
Lbp;1 .
7
7
5
1
Lbp;N

R7 X
R6
X

(17)
pU1  qU2 CB4

p1 q1

KLTI
(11)
Eq. (10) then becomes
b

Kz F  f B  f m  f m

(12)

By using these derivations, the ordinary differential equations of


motion in Eq. (1) are transformed into the linear algebraic equations in Eq. (12). The global Newton method can be used to calculate solutions of Eq. (12). The Jacobian matrix J for the global
Newton iteration is dened as



d
b
v Kz f B f m f m  F
(13)

vz

Smoothing functions are employed to approximate the piecewise nonlinear forces so that the explicit formulation of J can be
derived. These nonlinear forces include bearing reaction forces and
tooth loads on the drive-side and back-side of the gears. The inverse
Fourier transformation operator H is dened such that z Hz,
which maps from the time domain to the frequency domain.

2
6
H6
4

n
H1 H2

R4 P
R3
P
m1 l1

H3 m; l; R5

R7 P
R6
P
p1 q1

7
7
5

H4 p; q; R8
n

(14)
where H1eH4 are the inverse Fourier transformation operators
associated with G1eG4, respectively.
The explicit formation of J is obtained as

J KH

d
vf m

The Jacobian matrix associated with carrier-bearing forces, JB, is

pU1  qU2 2 MA4 U1 B1 U2 B2

3
7
Lm;b
r2;n 7

7
m;b 7
Lm;b
Lm;b
s2;1 / Ls2;n 7
s1;i
7
7
7
Lm;b
7
p;1
7
7
5
1

p1 q1
R4 X
R3
X

(16)

m1 l1

6
6
6
6
6
vf m
G6
H
6
6
vz
6
6
6
4 Symmetric

Lm;b
p;n

where KLTI is the time-invariant stiffness matrix dened in


Appendix B. The function G is not singular, so it can be eliminated
from both sides of Eq. (10). By dening

R7 X
R6
X

pU1  qU2 CB4 KLTI 5z

p1 q1

49

d
Hvf m =vzG

b
vf m

vf B
GH
GH
G
vz
vz
vz

(15)

Individual components in Jbm are dened in Appendix B.


J can be approximated numerically by employing nite difference algorithms (Turner and Walker, 1992). The numerical determination of J is time consuming in high dimensional space.
Furthermore, its accuracy depends on the step size of the nite
difference algorithms and can be disrupted by computational
round-off errors. Using the line search technique (Grippo et al.,
1986) for the Newton method improves the iteration convergence
rate.
4.1. Arc-length continuation and stability analysis
Arc-length continuation traces coexisting stable and unstable
solutions of the dynamic response. The direction of the estimated
next solution is selected to be perpendicular to the direction of the
previous solution. The iteration convergence rate relies on the
initial guess and the step size in the arc-length direction.
Solution stability is determined using FloqueteLiapunov theory
(Nayfeh and Balachandran, 1995). A perturbation Dz to the calculated periodic response z is introduced. The equations of motion are
linearized for small Dz to give

Dz0
Dz00


 0
 Jt  K

I
C

Dz
Dz0


(18)

The monodromy matrix of the linear system (Eq. (18) is computed


numerically by integrating the transformation matrices in multiple
time intervals (Kahraman, 1992; Friedmann, 1990). Eigenvalues of
the monodromy matrix are Floquet multipliers that characterize
the solution stability and bifurcation.
5. Model validations
Experiments on PG-A provide a benchmark for the lumpedparameter model using the extended harmonic balance method;
however, the relevant available experimental data for PG-A is only
at a single speed. Thus, results from the extended harmonic balance

50

Y. Guo et al. / European Journal of Mechanics A/Solids 47 (2014) 45e57

Fig. 2. (a) The lumped-parameter model and (b) side view of the planetary gear set.

method are also compared to the time-integrated results of the


lumped-parameter model of PG-B within the speed range of 0e
200 Hz.
5.1. Comparison to experimental data
The GRC project instrumented two identical 750-kW wind turbine gearboxes for the dynamometer (Fig. 3) and eld testing. Internal measurements include gear tooth loads, main shaft torque
and bending, internal component deections and misalignments,
and planet-bearing loads (Link et al., 2011).
The dynamic response of PG-A is computed using the extended
harmonic balance method at rated speed and rated torque. 5%
modal damping is assumed for PG-A based on the modal damping
extracted from a numerical torque impulse test of the PG-B nite
element model. Torque applied to the lumped-parameter model is
scaled to exclude the out-of-plane tooth loads caused by the planet
gear helix angle. The translational displacement of the carrier over
a carrier cycle computed using the extended harmonic balance
method is compared against the time-synchronized experimental
data measured in the GRC project as shown in Fig. 4. The agreement
between the extended harmonic balance method and experimental
data is reasonably good in capturing the maximum displacement
amplitude. Differences between the analytical model and experiments are present in the frequency domain.
The frequency spectrum of the carrier displacement (in Fig. 4) is
shown in Fig. 5. A modal component at 6P (six times the carrier
rotational frequency) is dominant in the frequency spectrum of the

measured signal. The passing frequency of the carrier planet pin


bores could be the main contributor to this 6P component. The
carrier is a bridged design and thus has three pin bores on both the
upwind and downwind sides. These two circular sets of pin bores
are misaligned because of manufacturing tolerances and torque
windup in operation, which could be the cause of the 6P excitation.
This 6P component might also be caused by the in-plane deformation of the carrier rim due to its varying thickness (Oyague et al.,
2010). The magnitude of the 6P component is of the 1P magnitude
for the calculated carrier displacement. The lumped-parameter
model considers the carrier as a single lumped mass; thus, it does
not consider the detailed carrier geometry. Consequently, the
lumped-parameter method underestimates the magnitude of the
displacement at 6P. The analytical results are dominated by 1P and
3P excitations due to the carrier rotation and planet symmetry. The
experiments include shaft misalignments, manufacturing tolerances, assembly errors, which cause high vibration amplitudes
compared to the analytical model that does not include these imperfections during the assembly and manufacturing. The experimental data includes more excitation sources than the analytical
model, resulting in the richer frequency spectra than that of the
model. In addition, the displacement of the rst mesh frequency
harmonic at 99P is more than an order of magnitude lower than the
displacements at the low-speed excitations (1Pe6P) for both the
calculated and measured results.
Calculated planet-bearing forces are compared against the
experimental data in Fig. 6. The mean and uctuating amplitudes of
the bearing forces computed by the extended harmonic balance

Fig. 3. A 750-kW wind turbine gearbox of the NREL GRC during a dynamometer test.

Y. Guo et al. / European Journal of Mechanics A/Solids 47 (2014) 45e57

Fig. 4. Carrier radial displacement with respect to the ring calculated using the
extended harmonic balance method (red e) and measured by the GRC project (/). (For
interpretation of the references to colour in this gure legend, the reader is referred to
the web version of this article.)

method largely match those of the experimental data. The planetbearing force for each individual planet bearing over a carrier cycle
is asymmetric because of different pin position errors and bearing
clearances for each planet. The lumped-parameter model predicts a
1.4% difference between the peak amplitudes of the planet-bearing
forces over a carrier cycle because of these errors and clearances.
The measured bearing force of planet 2 deviates from the other two
planets, which is caused by a signal offset error.
5.2. Comparison to computational methods
Dynamic responses computed by the extended harmonic balance method are compared against nite element and numerical
integration analyses for PG-B. In lightly damped systems, nite
element analysis and numerical integration require many time
steps for the transient response to diminish so that steady-state
data can be obtained. The extended harmonic balance method
avoids these long duration integration simulations by calculating
the steady-state dynamic response in the frequency domain. Numerical integration and nite element analysis compute only stable

51

Fig. 6. Measured (/) and calculated (red e) planet-bearing forces over a carrier cycle.
(For interpretation of the references to colour in this gure legend, the reader is
referred to the web version of this article.)

responses. Conversely, the harmonic balance method uses arclength continuation to nd unstable responses by tracing solutions as the speed changes quasi-statically.
The large ring gear mass results from a typical arrangement of
wind turbines whereby much of the gearbox mass is rigidly connected to the ring, which is supported on a relatively compliant
foundation. The input torque is applied to the carrier, and the sun
gear is the output. Gravity acts at the center of mass of all of the
component gears. For the ring and sun, it is a periodically varying
external excitation in the carrier reference frame in which the
model is formulated. One carrier period is analyzed using the nite
element model, giving 68 mesh cycles. Each mesh cycle is divided
evenly into 10 intervals. The nite element solutions have single
precision accuracy while the convergence tolerance is
100  1012 mm for analytical solutions. A global Newton iteration
scheme with line search technique (Baker and Overman, 2000) is
used to obtain accurate numerical solutions of the analytical model.
The line search technique makes the predicted solution at each
iteration always closer to the exact solution by adjusting the iteration step size for each step. For the numerical integration, 120
mesh cycles were simulated at each speed, of which 20 cycles are
considered as transient time. Each mesh cycle is discretized into 50
intervals. For the harmonic balance method, R1 R2 3, R3 1,
R4 4. Each carrier period is discretized into 5291 equally spaced
intervals. The solutions of the harmonic balance method have the
convergence tolerance of 1  106 mm.
The natural frequencies of PG-B below 400 Hz without bearing
clearance and nonlinear tooth contact are listed in Table 3. Modal
analysis is performed numerically on the nite element model of
PG-B by applying torque and force impulses at the carrier and sun.
Numerical impulse tests provide natural frequencies and their mode
shapes as described in Table 3. The natural frequencies predicted by
Table 3
Natural frequencies of PG-B.

Fig. 5. Frequency spectrum of the calculated (red e) and measured (/) carrier
displacement. (For interpretation of the references to colour in this gure legend, the
reader is referred to the web version of this article.)

Natural Frequency
Lumped-parameter
(Hz)

Natural Frequency Deviation Mode Type


Finite Element (Hz) (%)

Damping
(%)

40.75
69.49
157.53
301.46

40.94
69.1
163.7
312.1

5.48
5.26
7.43
5.19

0.35
0.56
3.77
3.41

Translational
Rotational
Translational
Rotational

52

Y. Guo et al. / European Journal of Mechanics A/Solids 47 (2014) 45e57

Fig. 7. Root-mean-square (RMS, mean removed) dynamic response of PG-B for various mesh frequencies calculated by the extended harmonic balance method (red C) and
numerical integration (/). Unstable solutions calculated by the harmonic balance method are denoted by (red B). (For interpretation of the references to colour in this gure
legend, the reader is referred to the web version of this article.)

Fig. 8. (a) Dynamic carrier-bearing force of PG-B calculated using the harmonic balance (red e) and numerical integration methods (e), (b) percentage of bearing contact in a carrier
cycle of PG-B with various speeds, given Dc 1 mm. (For interpretation of the references to colour in this gure legend, the reader is referred to the web version of this article.)

the lumped-parameter model match those obtained by the nite


element approach. The modal damping ratios are extracted from the
frequency response functions obtained from numerical impulse
tests of the nite element model, with damping ratios obtained from
the half-power points of the resonances listed in Table 3.
Quasi-static results of the lumped-parameter model have been
correlated with nite element analyses in Guo and Parker (2010) for
tooth loads and planet-bearing forces. As shown in Fig. 7, translational displacements of the ring and sun are compared for the
extended harmonic balance and numerical integration methods.
The agreement between these two methods is very good up to
100 Hz. The nonlinear resonance at 40 Hz is caused by tooth
wedging. Using speed sweeps, numerical integration predicts
nonlinear discontinuities in the dynamic response. Harmonic balance calculates the unstable branches between these nonlinear
jumps, denoted by the symbol (B). The agreement among the
proposed method, numerical integration, and nite element analysis validates this method for predicting nonlinear dynamic responses of wind turbine planetary gear sets.

A unique nonlinear effect caused by uctuating bearing contact


is present in the dynamic response as shown in Fig. 8(a). Good
agreement between the numerical integration and harmonic balance methods is evident. Two unstable branches are displayed in
the crossing resonance: curves BC and DE. Increasing the speed, the
response computed by numerical integration has the rst discontinuity at the point B and the second at the point D. Decreasing the
speed, the rst discontinuity to higher amplitude occurs at the
point E. Then the response has the rst discontinuity to lower
amplitude at the point C. The nonlinear effect displays multiple
types of bifurcation points including saddle-node bifurcation points
BeE and the transcritical bifurcation point A.
The bearing contact in a carrier cycle under the same conditions
of Fig. 8(a) is shown in Fig. 8(b). When the carrier speed increases
from 16 to 20 Hz, the percentage of bearing contact in a carrier cycle
decreases. The resonance bends to the low-frequency range at
point B, a softening effect. When the carrier speed decreases from
22 to 20 Hz, the bearing comes into contact. The response curve
bends to the high-frequency range at point E, a hardening effect.

Y. Guo et al. / European Journal of Mechanics A/Solids 47 (2014) 45e57

53

Fig. 9. (a) RMS dynamic response of PG-B carrier translational vibration with gravity (e) or mesh stiffness (e e) excitations. Bearing clearance is zero, innite backlash is used,
H harmonic, T translational mode, and R rotational mode; (b) RMS dynamic response of ring translational vibration when mesh stiffness is uctuating (e) and constant (e e).
The dotted lines denote the unstable branches. Dc 0.1, 0.5, 1 mm.

This resonance includes these two competitive nonlinear effects


simultaneously.
6. Dynamic response of planetary gear sets with gravity
Dynamic analyses of PG-A and PG-B were performed using the
extended harmonic balance approach. The response included 100
harmonics of the carrier frequency, three harmonics of the mesh
frequency, and eight upper and lower side bands of the mesh frequency harmonics. These parameters were selected based on the
energy distribution in the frequency spectrum of the experimental
data in Fig. 5.
Vibration modes of two-dimensional planetary gear sets include
rotational modes with distinct natural frequencies, translational
modes with degenerate natural frequencies of multiplicity two, and
planet modes with degenerate natural frequencies of multiplicity
N  3. Translational modes include translation but no rotation of

the carrier, ring, and sun gears. The rotational modes considered
have only rotation of the carrier, ring, and sun gears. The planet
modes considered have only planet motions; the carrier, ring, and
sun gears do not move. These vibration modes are derived from the
Eigenvalue problem of the equations of motion in Eq. (1) without
tooth contact and bearing clearance nonlinearities. The natural
frequencies of PG-A under 800 Hz are the translational modes at
165, 311, and 796 Hz and the rotational mode at 292 Hz. The natural
frequencies and mode shapes of PG-B are described in Table 3.
6.1. Gravity-induced excitation
The effects of gravity and uctuating mesh stiffness on the dynamic response of PG-B are shown in Fig. 9(a). Within the speed
range of interest, the carrier response with gravity as the excitation
source is one order of magnitude higher than that with uctuating
mesh stiffness. This agrees with the observation that low-frequency

Fig. 10. (a) Dynamic response of the RMS of carrier translational (upper graph) and rotational (lower graph) responses of PG-B with different backlash; (b) drive-side and back-side
tooth contact at the rst sun-planet mesh with one-third of nominal backlash.

54

Y. Guo et al. / European Journal of Mechanics A/Solids 47 (2014) 45e57

Fig. 11. RMS xr and x1 dynamic responses of PG-A with bearing clearance from 0 to 400 mm.

external excitations have a greater contribution to the dynamic


response than the high-frequency internal excitation from gear
meshing, as measured in the experiments shown in Fig. 5. In the
dynamic response, with gravity as the excitation source, only the
rst and second translational modes are present. In the dynamic
response, with uctuating mesh stiffness as the excitation, the vibration amplitude is much lower (although more resonances are
excited). The vibration resonances include the rst two translational modes, the fourth harmonic of the third translational
mode, and the second harmonic of the rotational mode. Fluctuating
mesh stiffness consists of the fundamental and higher harmonics of
the mesh frequency and can thus excite gearbox vibration modes
beyond the frequency range studied in Fig. 9(a).
Under gravity effects, the dynamic responses of ring translation
with and without mesh stiffness variation are compared in Fig. 9(b)
with various bearing clearance. The hardening effects shown in the
gure when bearing clearance is larger than 0.5 mm is caused by
bearing-raceway contacts. With increased bearing clearance, the
resonances shifts to lower speed because of the reduced bearing
stiffness, which is detailed in Section 6.2. When the mesh stiffness

variation is included, the vibration amplitude is higher than that


with constant mesh stiffness. However, the resonance shapes are
the same in nature. This indicates a mild coupling between gravity
and the mesh stiffness uctuation.
6.2. Nonlinear dynamic effects
Fig. 10(a) (top) and (bottom) shows the carrier translational and
rotational displacements of PG-B with different tooth backlash
conditions. Backlash affects the translational mode resonance at
41 Hz as shown in Fig. 10(a) (top graph), but it does not affect the
carrier rotational displacement as shown in Fig. 10(a) (bottom
graph). With nominal backlash (bs br 482 mm), the shape of the
resonance peak bends toward the right. This nonlinear effect is the
hardening effect caused by tooth wedging at the resonance. As the
backlash decreases, the shape of the resonance bends further to the
right and the resonant frequency increases from 41 to 84 Hz.
When the backlash is one-third of its nominal value, a closed
loop caused by nonlinear tooth contact occurs at 70 Hz in the dynamic response of xc. Fig. 10(b) shows the tooth contact at both the

Fig. 12. Back-side tooth load at the rst sun-planet mesh and planetary load sharing over a carrier cycle (a) with various bearing clearances and speeds; (b) with various sun
stiffnesses and bearing clearances for PG-A. The carrier speed is 0.5 Hz in (b).

Y. Guo et al. / European Journal of Mechanics A/Solids 47 (2014) 45e57

drive-side and back-side of the rst sun-planet mesh with onethird of the nominal backlash under the same condition as
Fig. 10(a). In this gure, a y-axis value of 1 indicates a full-contact
condition and a 0 represents an out-of-contact condition. Within
the speed range of 70e85 Hz, the gear rotates signicantly because
of the rotational mode at 70 Hz and loses tooth contact. In the
meantime, the translational mode causes the radial motion between the sun and planet 1. This radial motion exceeds the available
tooth radial gap, which leads to tooth wedging. The coexistence of
tooth wedging and tooth contact loss within the same speed range
reects the competition between the translational and rotational
modes, resulting in the closed loop in Fig. 10(a).
Fig. 11(a) and (b) shows the dynamic responses of xr and x1, with
various values of bearing clearance for PG-A. The resonant frequency of the rst translational mode decreases from 165 to 28 Hz
when the bearing clearance increases from 0 to 400 mm as shown in
Fig. 11(a). Bearing clearance affects the resonant frequencies of the
translational mode shapes of the member with the bearing clearance. For large-scale wind turbine gearboxes, the resonant frequencies can be further reduced into the range of external lowspeed excitations because of structural exibilities of the drivetrain, the blades, and the supporting components (Helsen et al.,
2010). The responses without clearance and with clearance of
400 mm, which is essentially innite clearance, correspond to two
limiting systems. The mode shapes of these two limiting systems
are different from each other as depicted in Fig. 11(a). In Fig. 11(a),
the ring gear displacement increases 4.6 times when clearance
increases from 0 to 400 mm. In the meantime, the planet translational displacement x1 decreases 25 mm when the bearing clearance increases. Clearance in the carrier bearing increases the
vibration of the central members: the sun, ring, and carrier, but its
effect on planet responses is much smaller. The bearing forces at the
carrier support are small because they are nearly balanced by corresponding self-imposed tooth loads. Because of the absence of the

0
60
Gl a; b; c 6
4
0

cosaU1  bU2 s1
cosaU1  bU2 s2

cosaU1  bU2 sn

sinaU1  bU2 s1
sinaU1  bU2 s2

sinaU1  bU2 sn

/
/
/
/

bearing clearance is larger than the backlash, the back-side tooth


load and load sharing factor gradually increase with bearing
clearance. When the speed is near the rpm that excites the resonant
frequency, the tooth load and load sharing factor are higher than
when the excitation speed is further away from the resonant frequency. The critical clearance when tooth wedging occurs is
entirely independent of speed. The critical clearance value is
reached when the bearing clearance equals the backlash. Therefore,
the bearing clearance should be smaller than the backlash to avoid
the potential of tooth wedging in wind turbine planetary gear sets.
The sun support stiffness also affects tooth wedging. Fig. 12(b)
shows the back-side tooth load over a carrier cycle at various sun
support stiffness values. When the bearing clearance equals or is
larger than the backlash, the tooth load rises exponentially with an
increase in sun support stiffness. At zero sun support stiffness, no
back-side tooth contact occurs, regardless of the value of bearing
clearance. When the bearing clearance is smaller than the backlash,
the back-side tooth load is zero for small sun support stiffness
values. However, when the sun support stiffness has the same order
of magnitude as the ring stiffness, tooth wedging occurs even with
a smaller bearing clearance than the backlash.
7. Conclusions
A harmonic balance method was developed to calculate the
dynamic response of wind turbine planetary gear sets. Results
obtained compare well with the GRC experiments of PG-A and the
numerically integrated results of the lumped-parameter model of
PG-B. This approach avoids protracted time integration simulations
by calculating the dynamic response in the frequency domain. The
time savings of this method make it suitable for parametric studies,
which could be used to tune the planetary gear dynamics during
the design phase to minimize vibratory response.
Gravity introduces fundamental excitations in the rotating car-

sin caU1  bU2 s1


sin caU1  bU2 s2

sin caU1  bU2 sn

nominal bearing force, the carrier-bearing clearance creates


nonlinear dynamic behavior. Away from the resonance, the vibration amplitude is too low, so the bearing is not in contact. Near the
resonance, the carrier-bearing rolling element displacement
evolves gradually from no contact to partial contact, and eventually
to full contact at the resonant frequency. The carrier-bearing rollers
contacting the raceways at the resonant frequency induces the
stiffening effects shown in Fig. 11(a) and (b). This stiffening effect
reects the increase in the overall bearing stiffness from the
increased bearing contact (Guo and Parker, 2012a).
6.3. Tooth wedging
Tooth wedging induced by gravity is a potential source for
planet-bearing premature failures (Guo and Parker, 2010; Larsen
et al., 2003; Rasmussen et al., 2004; Hansen et al., 2011). The inuences of bearing clearance on tooth wedging and planetary load
sharing are investigated in Fig. 12(a). The tooth loads are calculated
for various values of the ratio between bearing clearance and
backlash. When the bearing clearance is smaller than the backlash,
there is no back-side tooth contact regardless of speed. When the

55

3
0
07
7
5
0

(20)

rier frame of planetary gear sets. It plays a dominant role in the


dynamics of wind turbine planetary gear sets compared to excitations caused by mesh stiffness uctuation. Gravity causes tooth
wedging and bearing contact at vibratory resonances, leading to
stiffening effects in the dynamic response. A unique nonlinear effect induced by gravity illustrates the interaction of tooth wedging
and tooth contact loss in the translational and rotational modes.
Clearance in the carrier bearings reduces the resonant frequencies of the translational modes of the carrier, which are the
lowest resonant frequencies of the examined systems. The shapes
of these translational modes are affected by this bearing clearance.
For wind turbine gearboxes larger than the ones studied, these
frequencies could be reduced into the range of external low-speed
excitations.
Tooth wedging disrupts planetary load sharing and reduces the
life of planet bearings. However, it can be prevented if the carrierbearing clearance is less than the tooth backlash.

56

Y. Guo et al. / European Journal of Mechanics A/Solids 47 (2014) 45e57

Acknowledgment

This work was supported by the U.S. Department of Energy


under Contract No. DE-AC36-08GO28308 with the National
Renewable Energy Laboratory.

3
Xcbrj sin jbri sin ar Xcbrj sin jbrj cos ar Xcbrj sin jbrj
6
7
Lm;b
4 Xcbrj cos jbrj sin ar Xcbrj cos jbrj cos ar Xcbrj cos jbrj 5
r2;j

Xcbrj sin ar

Xcbrj cos ar

Xcbrj

(27)
2

Xcbsj sin jbsj cos jbsj Xcbsj sin jbsj


6
b
b
b7
b
b
2 b
b
7
6
Lm;b
4 Xcsj sin jsj cos jsj Xcsj cos jsj Xcsj cos jsj 5
s1;j
b
b
b
b
b
Xcsj sin jsj
Xcsj cos jsj
Xcsj

Appendix A. Operators for the harmonic balance approach

1 cos Ui s1 sin Ui s1 cos 2Ui s1


6 1 cos Ui s2 sin Ui s2 cos 2Ui s2
Gi 6
4

1 cos U1 tn sin U1 sn cos 2U1 sn

3
/ sin Ri Ui s1 0
/ sin Ri Ui s2 0 7
7
/

5
/ sin Ri U1 sn 0

Xcbsj sin2 jbsj

(28)
3

(19)

Xcbsj sin jbsj sin as Xcbsj sin jbsj cos as Xcbsj sin jbsj
7
6
Lm;b
4 Xcbsj cos jbsj sin as Xcbsj cos jbsj cos as Xcbsj cos jbsj 5
s2;j
b
b
b
Xcsj cos as
Xcsj sin as
Xcsj
(29)

where i 1,2.
where l 3, 4, plus or minus signs are used in the operator,
respectively. When l 3, 4, c R5, R8, respectively.

Ai 4

3
A0i

Xcbrj sin ar cos ar Xcbrj sin ar 3


Lm;b
4 Xcbrj sin ar cos ar
Xcbrj cos2 ar
Xcbrj cos ar 5
p;j
2

Xcbrj sin2 ar
Xcbrj sin ar

(21)

B 4

1
2

3
B0i

0
6 1
6
6
6
0
Bi 6
6
6
6
4

5
n

0
2

2
0
1
0
Xi

7
7
7
7
7
7
7
7
Xi 5
0

1
2

(31)

(24)

where drj ; dsj denote the back-side mesh deection vectors at the
jth ring-planet and sun-planet meshes. Parameter s controls how
close the smoothing functions are to the original piecewise
nonlinear functions.

(32)

 b
s
L

 b
dbrj G zyr cos jbrj zxr sin jbrj zzi sin ar zhi cos ar zui zur  r
L

dbsj G zys cos jbsj zxs sin jbsj zzi sin as zhi cos as zui zus 

jbsj jj as ; jbrj jj  ar
(33)

(25)

where jj denotes the position angle of the jth planet. ar, as denote
the pressure angles of the ring and sun gear teeth. br, bs denote the
tooth backlashes of the ring and sun gears.
References

where P diag(kl,u, ., kl,u), l c,r,s.


By dening X() Hdiag()G,

cbsj  kbsj t s sec h2 sdbsj dbsj 1 tan hsdbsj

Appendix B. Jacobian matrix for the harmonic balance


approach

KLTI diag0; 0; Pc ; 0; 0; Pr ; 0; 0; Ps ; 0; .; 0

(23)
3

1
0

Xcbsj

cbrj  kbrj t s sec h2 sdbrj dbrj 1 tan hsdbrj

where X1 R1, X2 R2, X3 R5, X4 R8.

Xcbsj cos as

(30)
(22)

Xcbrj

2 Xcb sin2 a
Xcbsj sin as cos as Xcbsj sin as 3
s
sj
Xcbsj cos as 5
4 Xcbsj sin as cos as Xcbsj cos2 as
Xcbsj sin as



A0i diag 0; 12 ; 12 ; 22 ; 22 ; .; Xi2 ; Xi2 ; 0 ; i 1; 2; 3; 4

Xcbrj cos ar

Xcbrj sin2 jbrj Xcbrj sin jbrj cos jbrj Xcbrj sin jbrj
7
6
b
b
b7
b
b
2 b
b
Lm;b
6
4 Xcrj sin jrj cos jrj Xcrj cos jrj Xcrj cos jrj 5
r1;j
Xcbrj sin jbrj
Xcbrj cos jbrj
Xcbrj
(26)

Al-shyyab, A., Kahraman, A., 2005a. Non-linear dynamic analysis of a multi-mesh


gear train using multi-term harmonic balance method: period-one motions.
J. Sound Vib. 284 (1e2), 151e172.
Al-shyyab, A., Kahraman, A., 2005b. Non-linear dynamic analysis of a multi-mesh
gear train using multi-term harmonic balance method: subharmonic motions.
J. Sound Vib. 279 (1e2), 417e451.
Ambarisha, V.K., Parker, R.G., 2007. Nonlinear dynamics of planetary gears using
analytical and nite element models. J. Sound Vib. 302, 577e595.
Bahk, C.-J., Parker, R.G., 2011. Analytical solution for the nonlinear dynamics of
planetary gears. ASME J. Comput. Nonlinear Dynam. 6 (2), 021007.
Baker, G., Overman, E., 2000. The Art of Scientic Computing. William H. Press.

Y. Guo et al. / European Journal of Mechanics A/Solids 47 (2014) 45e57


Blankenship, G.W., Kahraman, A., 1996. Gear dynamics experiments. Part-I. Characterization of forced response. In: ASME Power Transmission and Gearing
Conference, San Diego.
Botman, M., 1976. Epicyclic gear vibrations. J. Eng. Ind. 97, 811e815.
Friedmann, P.P., 1990. Numerical methods for the treatment of periodic systems
with applications to structural dynamics and helicopter rotor dynamics. Comput. Struct. 35 (4), 329e347.
Grippo, L., Lampariello, F., Lucidi, S., 1986. A nonmonotone line search technique for
Newtons method. SIAM J. Numer. Anal. 23 (4), 707e716.
Guo, Y., Parker, R.G., 2010. Dynamic modeling and analysis of a spur planetary gear
involving tooth wedging and bearing clearance nonlinearity. Eur. J. Mech. A
Solids 29, 1022e1033.
Guo, Y., Parker, R.G., 2012a. Dynamic analysis of planetary gears with bearing
clearance. ASME J. Comput. Nonlinear Dynam. 7 (3), 031008.
Guo, Y., Parker, R.G., 2012b. Stiffness matrix calculation of rolling element bearings
using a nite element/contact mechanics model. Mech. Mach. Theor. 51, 32e45.
Guo, Y., Keller, J., LaCava, W., 2012. Combined effects of input torque, non-torque
load, gravity, and bearing clearance on planetary gear load share in wind turbine drivetrains. In: AGMA Fall Technical Meeting.
Gurkan, N. E., Ozguven, H. N., 2007. Interactions between backlash and bearing
clearance nonlinearity in geared exible rotors. In: ASME International Design
Engineering Technical Conferences DETC2007-34101.
Hansen, A.M., Rasmussen, F., Larsen, T.J., 2011. Gearbox Loads caused by Double
Contact Simulated with HAWC2. European Wind Energy Association.
Helsen, J., Vanhollebeke, F., Coninck, F.D., Vandepitte, D., Desmet, W., 2010. Insights
in wind turbine drive train dynamics gathered by validating advanced models
on a newly developed 13.2 MW dynamically controlled test-rig. Mechatronics
21 (4), 737e752.
Kahraman, A., 1992. On the response of a preloaded dynamic oscillator with a
clearance: period-doubling and chaos. Nonlinear Dynam. 3, 183e198.
Kahraman, A., Blankenship, G. W., 1994. Planet mesh phasing in epicyclic gear sets.
In: International Gearing Conference, Newcastle, UK, pp. 99e104.
Kahraman, A., Singh, R., 1991. Interactions between time-varying mesh stiffness and
clearance non-linearities in a geared system. J. Sound Vib. 146 (1), 135e156.
Kahraman, A., Vijayakar, S.M., 2001. Effect of internal gear exibility on the quasistatic behavior of a planetary gear set. J. Mech. Des. 123 (3), 408e415.
Larsen, T.J., Thomsen, K., Rasmussen, F., 2003. Dynamics of a Wind Turbine Planetary Gear Stage. Ris National Laboratory, Ris-I-2112 (EN).

57

Link, H., LaCava, W., Van Dam, J., McNiff, B., Sheng, S., Wallen, R., McDade, M.,
Lambert, S., Buttereld, S., Oyague, F., 2011. reportGearbox Reliability Collaborative Project Report: Findings from Phase 1 and Phase 2 Testing. National
Renewable Energy Laboratory.
Musial, W., Buttereld, S., McNiff, B., 2007. Improving wind turbine gearbox reliability. In: European Wind Energy Conference, Milan, Italy, NREL/CPe500e
41548.
Nayfeh, A., Balachandran, B., 1995. Applied Nonlinear Dynamics. John Wiley and
Sons.
Oyague, F., Gorman, D., Sheng, S., 2010. NREL Gearbox Reliability Collaborative
Experimental Data Overview and Analysis. NREL.
Parker, R.G., Agashe, V., Vijayakar, S.M., Sep. 2000a. Dynamic response of a planetary
gear system using a nite element/contact mechanics model. J. Mech. Des. 122
(3), 304e310.
Parker, R.G., Vijayakar, S.M., Imajo, T., Oct. 2000b. Non-linear dynamic response of a
spur gear pair: modelling and experimental comparisons. J. Sound Vib. 237 (3),
435e455.
Raghothama, A., Narayanan, S., 1999. Bifurcation and chaos in geared rotor bearing
system by incremental harmonic balance method. J. Sound Vib. 226 (3), 469e492.
Rasmussen, F., Thomsen, K., Larsen, T., 2004. The Gearbox Problem Revisited. Ris
National Laboratory, AED-RB-17 (EN).
Seydel, R., 1994. Practical Bifurcation and Stability Analysis: From Equilibrium to
Chaos, second ed., Springer.
Singh, A., 2010. Load sharing behavior in epicyclic gears: physical explanation and
generalized formulation. Mech. Mach. Theor. 45, 511e530.
Singh, A., 2011. Epicyclic load sharing map e development and validation. Mech.
Mach. Theor. 46, 632e646.
Thomsen, J., 2003. Vibration and Stability: Advanced Theory, Analysis, and Tools,
second ed., Springer.
Turner, K., Walker, H.F., 1992. Efcient high accuracy solutions with gmres(m). J. Sci.
Stat. Comput. 13 (3), 815e825.
Velex, P., Flamand, L., Mar. 1996. Dynamic response of planetary trains to mesh
parametric excitations. J. Mech. Des. 118 (1), 7e14.
Vijayakar, S.M., 1991. A combined surface integral and nite element solution for a
three-dimensional contact problem. Int. J. Numer. Methods Eng. 31, 524e546.
Zhu, F., Parker, R.G., Jan. 2005. Non-linear dynamics of a one-way clutch in belt
pulley systems. J. Sound Vib. 279 (1e2), 285e308.

Das könnte Ihnen auch gefallen