Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Mechanism and Machine Theory 155 (2021) 104122

Contents lists available at ScienceDirect

Mechanism and Machine Theory


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

Research paper

Modeling and analysis of pure kinematic error in harmonic


drive
Hang Jia a,b, Junyang Li a,b,∗, Guo Xiang a,b, Jiaxu Wang a,b, Ke Xiao a,b,
Yanfeng Han a,b
a
College of Mechanical Engineering, Chongqing University, Chongqing 40044, China
b
State Key Laboratory of Mechanical Transmissions, Chongqing University, Chongqing 40044, China

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

Article history: Harmonic drives are widely used in the precision positioning systems due to their at-
Received 7 May 2020 tractive attributes. However, kinematic error is the most fundamental problems in these
Revised 30 August 2020
drives. In this work, we present a method for analyzing the pure kinematic error of har-
Accepted 24 September 2020
monic drive with machining and assembly errors. The pure kinematic error is decomposed
into two parts including nonuniform motion component and lag component. The beat fre-
Keywords: quency phenomenon of the pure kinematic error is revealed. The explanation of the mech-
Harmonic drive anism of pure kinematic error is verified by the experimental data from a specialized har-
Kinematic error monic reducer test device. The influences of the wave generator structure (automatic align-
Machining errors ment and integral structures) and shape on the pure kinematic error of harmonic drive are
Assembly errors discussed and quantitatively evaluated, and the one possible new configuration of the wave
generator is presented, which can improve the transmission performance of the drive. The
results of this paper are believed to be useful for the design of harmonic drive.
© 2020 Published by Elsevier Ltd.

1. Introduction

The harmonic drive (HD) is well known for its high reduction ratio, low weight, small volume, and high efficiency [1],
and it is used widely in the field of robotics [2]. These may benefit from its unique transmission principle and simple geo-
metric structure. A HD contains only three main components as shown in Fig. 1. The wave generator (WG) is an elliptically
shaped steel core surrounded by a flexible race bearing. The flexspline (FS) is a thin-walled flexible cup having two fewer
teeth on its outer rim than on the inner rim of the circular spline (CS). The CS is a rigid steel ring with teeth machined
into the inner circumference. Understanding the mechanism of kinematic error is expected to improve the control perfor-
mance of harmonic drive-based devices such as robot manipulators [3], and the transmission performance of HD from a
design perspective purposefully. Many studies have been conducted on HD kinematic error [4–16], nonlinear stiffness [17],
hysteresis [18,19], and dynamic friction [20]. However, the kinematic error often showed significant variation with assembly
conditions and was difficult to characterize accurately, and a seemingly similar HD can produce different kinematic error
signatures [21,22]. As researcher Bikash [23] pointed out that kinematic error from gearing laws in HDs and relevant issues
are still remained largely unexplained.


Corresponding author at: College of Mechanical Engineering, Chongqing University, Chongqing 40044, China.
E-mail address: lijunyang1982@sina.com (J. Li).

https://doi.org/10.1016/j.mechmachtheory.2020.104122
0094-114X/© 2020 Published by Elsevier Ltd.
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 1. Harmonic drive components.

Given its importance, many investigations have focused on kinematic error in HD. Emelyanov [4] made a mathematical
analysis for the origin of the kinematic error and showed that the kinematic error is caused by assembly and machining
imperfections on the three principal components of HD. He found that the error is periodic and its main frequency is twice
that of the WG rotation. Tuttle and Seering [5] carried out extensive experiment to analyze kinematic error of the HD.
Their understanding of the source of the kinematic error is consistent with that of Emelyanov [4]. However, Hsia [6] and
Ramson [7] both proposed that the kinematic error in HD is caused by the deformation of the FS. In addition, Nye and
Kraml [8] classified the kinematic error into two categories, which are caused by internal and external factors respectively,
and found that the kinematic error is dynamic and uncertain. However, the above analysis of kinematic error is based on
phenomenological analysis method. In order to improve the transmission precision of HD from a design perspective, it is
necessary to analyze the mechanism of kinematic error.
Hsia [6] first proposed the numerical formulation of the kinematic error based on elliptic WG. He argued that the kine-
matic error is the inherent property of HD operating principle. Based on the same kinematic error mechanism, Gravagno
et al. [9] further analyzed the influence of WG shape on the pure kinematic error. However, the compliant FS cup is ignored
in their derivation. Dong et al. [10] proposed a kinematic model of HD considering input eccentricity error, but the influence
of input eccentricity error on load distribution and backlash distribution was studied, and the effect of it on kinematic error
was not analyzed. In order to consider the influence of non-linear hysteresis on kinematic error, the kinematic error was
handled as the sum of the synchronous component and the nonlinear elastic component by Yamamoto et al. [11]. Tuttle
et al. [12] proposed that the kinematic error in HDs is caused mainly by gear-tooth manufacturing errors in both the CS and
FS, which was modeled as a sum of sinusoidal functions. In [13], the kinematic error was modeled using a sum of sinusoidal
functions to estimate torque for robotic joint with HD. The error was investigated by Ghorbel et al. [14] as two components,
the pure kinematic error and the torsional flexibility induced component, in which a fifteen-term Fourier series expansion
was used to fit the pure kinematic error experimental data. León et al. [15] investigated the influence of gear tooth ge-
ometrical variations on kinematic error with stochastic analysis. Assuming that the clearance size is invariant, Liu et al.
[16] studied the dependence of the kinematic error on clearance by translating the clearance of different positions into the
clearance between the harmonic gear teeth. However, no comprehensive understanding of the kinematic error mechanism
of HD with machining errors and assembly errors is available in these literatures.
For the purpose of definite understanding of the mechanism of kinematic error in HD, the pure kinematic error proposed
by Ghorbel et al. [14] is modeled and analyzed in this study. The influences of machining error, assembly error, and WG
shape and structure on pure kinematic error are discussed. A specialized harmonic drive test device is built and experiments
are carried out to verify the proposed model.
This paper mainly contains the following five parts. The first part is kinematic analysis of HD, then introduces the pure
kinematic error of HD and it is decomposed into a nonuniform motion component and a lag component. The next step is to
model the machining and assembly errors. In the next part, the pure kinematic error is analyzed and calculated theoretically
by modeling. Next, the influences of the WG shape and structure on the pure kinematic error of HD are discussed and
quantitatively evaluated. Finally, a specialized harmonic drive test device and test methods are introduced, and the proposed
model is demonstrated using data of tests.

2. Transmission principle and the modeling of pure kinematic error

2.1. Harmonic drive kinematic model

In a conventional HD kinematic model, a neutral line is defined as the intersection of the central surface that move
through the middle of FS wall thickness and a plane perpendicular to rotation axis. Other characteristic curves of FS are
the equidistance curve of the neutral line [9]. The neutral curve is assumed to maintain the same length (no strain) before

2
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 2. Kinematic model for the flexspline.

and after deformation [24], and there is no deformation at the closed (output) end of the FS cup [25]. According to the
neutral line no strain assumption, the neutral curves are constant velocity curves with the same speed at the deformed
and undeformed ends of the FS cup. As shown in Fig. 2, the angular velocity of a point on the neutral curve at the elastic
deformation end depends on the radius connecting the center of revolution with such point, and the following condition
must hold
 luc /vF S
ω (t )dt = 2π (1)
0
where luc is the total length of the neutral curve, and vFS is the constant velocity of a generic point on the curve. The
constant angular velocity of a generic point on the neutral curve at close end of the cup is focused in our analysis, since it
is the input or output speed of FS in HD. The angular velocity ωFS can be described as
vF S 2π vF S
ωF S = = (2)
ρ luc
According to the kinematic model of the FS shown in Fig. 2, the kinematic relationship between two ends of FS can be
described by the following equation
 luc /vF S
ω (t )dt
0
= ωF S (3)
luc /vF S
Focusing on the elastic deformation end of the FS, motion and energy transfers are executed through teeth meshing
between the FS and CS. Let ωWG be the angular velocity of WG (CS is fixed), the instantaneous transmission ratio yields
ωi
i= (4)
ωWG
where ωi is the angular velocity of a generic point on the neutral curve at the elastic deformation end of FS. One may
observe that the instantaneous transmission ratio is related to profile of the neutral curve. However, for the angular velocity
at the close end of the FS concerned in this paper, the FS has an integral average effect on the angular velocity at the elastic
deformation end as described in Eq. (3). In other words, the nonuniform angular velocity at the elastic deformation end is
converted to a single output to realize the desired uniform transmission ratio. Therefore, the profile of neutral curve (or WG
shape) has no effect on the kinematic error of HD in an ideal situation.
When the WG is inserted into the FS, the FS is deformed and the FS teeth and the CS teeth are meshed at the two sides
of major axis of the WG. In order to capture the fundamental kinematic principle of HDs, the FS of a HD is assumed to be a
ring without the cup, as shown in Fig. 3(a). But the kinematic relationship of the FS is described in Eq. (3). This assumption
is considered reasonable according to the studies of Dong et al. [24] and Chen et al. [26]. The main mechanism of the HD,
which is a two-degree-of-freedom system shown in Fig. 3(b), is analogous to a planetary gear train.
The ideal kinematic relationship of the three ports of the HD, applying the Willis formula to HDs, can be expressed as
follows
ωF S −ωWG ZCS
= (5)
ωCS −ωWG ZF S
where ωCS denote the angular velocity of CS, ZCS and ZFS denote the number of teeth of CS and FS, respectively. The most
common configuration for the HD consists of the WG as the input port, the FS as the output port, and the CS fixed to ground

3
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 3. Kinematic model of harmonic drive: (a) Planar model of harmonic drive, (b) Kinematic representation of a harmonic drive showing the three ports.

and held immobile. If we integrate Eq. (5) with respect to time and set the initial values to zero, the basic harmonic-drive
equations which capture the three-port behavior are
ZF S
θWG = − θF S = −NθF S (6)
ZCS − ZF S
where N is the gear reduction ratio, θ WG and θ FS denote the angular position of WG and FS, respectively. However, ex-
perimental observations show that the output is not linearly related to the input [21]. The causes of this nonlinearity are
torsional compliance in the HD components and the kinematic error that is due to machining and assembly errors in the
transmission. In a comprehensive model, the kinematic relationship can be expressed as follows [12]
θWG = −N (θF S +θe + θ ) (7)
where θ e is the kinematic error, θ is the torsional angle of the HD. The kinematic error is defined as the difference
between theoretical and actual output angles, which can be represented by the following equation [14]
∼ θm
θ = N − θl (8)

where θ m is the input angle attached to the WG, θ l is the rotation angle of output shaft connected to the FS. Combining
Eqs. (7) and (8), the kinematic error can be determined as

θ =θe + θ (9)
Eq. (9) indicates that the torsional angle due to torsional compliance in the HD is a component of the kinematic error.
The kinematic error can be divided into static and dynamic kinematic error [27]. For the kinematic error of HD, in [14],
it was decomposed into a pure component and a flexibility-induced component correspondingly. In this paper, significant
emphasis has been placed on the pure kinematic error in consideration of machining and assembly errors.

2.2. Composition of the pure kinematic error

The pure kinematic error is considered as a basic component of kinematic error, which results from the kinematic struc-
ture of the HD [14]. Clearly, any deviation of the teeth from an ideal shape results in the well-known errors in gear ratio

4
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 4. Pure kinematic error.

that is named the kinematic error. Emelyanov [4] and Tuttle and Seering [5] concluded that the error has a fundamental
component occurring at twice the WG rotation, and colored with higher frequency components. In [14], a Fourier series
expansion was used to fit the pure kinematic error experimental data.

a0 
k
θ˜p = + [an cos (nθm ) + bn sin (nθm )] (10)
2
n=1

Here, constants an and bn are determined from experimental data by numerical integration.
In this study we present a more systematic explanation from the viewpoint of mechanism of error and propose that the
pure kinematic error mainly depends on the following two parts. The first part, as the constant part in Eq. (10), is lag com-
ponent that mainly depends on the HD nonlinear torsional rigidity and clearance. For the purpose of tracking and locating,
the HD in the harmonic drive-based devices need to reverse frequently. In this case, the lag component plays an impor-
tant role in the transmission accuracy of the device. Moreover, considering the time-varying backlash due to misalignment
and tooth error, the lag component is not a constant. The second part is nonuniform motion component that depends on
the misalignment and tooth error caused by assembly and machining imperfections. As indicated in Eq. (10), the nonuni-
form motion component is usually captured using a superposition of sinusoids. Consequently, the expression of the pure
kinematic error can actually be decomposed into

θ˜p = θme + θle (11)


where θ me and θ le denote the nonuniform motion and lag components of pure kinematic error, respectively.
The following assumption is made for deriving the pure kinematic error due to machining and assembly errors for the
HD. Considering that the FS is a flexible body, the nonuniform motion component is assumed to be mainly caused by the
meshing tooth with the maximum meshing depth in meshing area, and the influence of other meshing teeth on the error
can be expressed by the error homogenization coefficient [28,29]. This assumption is supported by the results presented by
Ghorbel et al. [14]. The nonuniform motion component can be determined as

θme = ke ϑi (12)
where ϑi is kinematic error caused by single tooth engagement, and ke is error homogenization coefficient. In [28,29], the
error homogenization coefficient is given by the following expression
k
ke = B (13)
Zmp
where Zmp is the number of meshing teeth in one meshing area, and the value of coefficient kB is in the range 0.8–1.0.
This indicates that, with the existence of machining and assembly errors, WG shape can influence the kinematic error by
affecting the number of meshing teeth. Under an external load, neighboring teeth on the FS and the CS may be forced to
contact. In this paper, the relationship between the number of meshing teeth and the external load is not discussed.
As shown in Fig. 4, let B and C be two points fixed to the pitch curve of FS and CS respectively. Initially, B and C
are coincident, after the WG has rotated clockwise through an angle, ideally, B and C end up being at B’ and C’. The arc
lengths traversed by B and C are equal. In practice, point C end up being at point C” due to machining and assembly errors.

5
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 5. The definition of eccentricity and deflection errors of WG.

According to the kinematic model for the FS, we can find θ me by equating the length of the arc on the FS close end (it
does not physically exist) and on the FS deformation end (see Fig. 4 lC’C” = lDE ). The arc length lAD is equivalent to torsional
flexibility and clearance on the close end of FS.

lAD + lDE
θ˜p = 2π (14)
lFS
where lFS is the total length of the FS pitch curve.

3. Model method for harmonic drive with error

There are always random errors in the assembly and machining processes of HD, and these errors that lead to teeth
placement error have great influence on the pure kinematic error of the gear system. They are mainly divided into the
following two categories: the eccentricity and deflection errors of WG, FS and CS teeth errors.

3.1. Mathematical model of eccentricity and deflection errors

Following the planar kinematic model of HD, in this analysis the WG misalignment should be considered in the plane
xoy shown in Fig. 3(a). The points of the FS reference circle undergo a displacement when the WG is inserted into the FS.
And the inner wall of the FS is considered to be in tight contact with the outside edge of the WG. Therefore, WG and FS
have the same geometric center. Since the CS is rigid, the CS reference circle remains unchanged. As shown in Fig. 3(a), the
geometric centers of the FS and CS coincide with the rotation center O (input shaft) under the ideal assembly relation.
mZCS
RCS = LF = (15)
2
where m denotes tooth module.
In practice, the eccentricity and deflection errors of WG cannot be canceled due to the assembly and machining im-
perfections. Fig. 5 shows the schematic diagram of the errors for HD, where OC Z1 and OF Z3 are the shaft axis of CS and
WG, respectively, OZ2 is the input axis, and point OF is the geometric center of WG. The Y3 -axis of the coordinate system
OF -X3 Y3 Z3 is in coincidence with the major axis of WG, and the planes X3 OF Y3 and X1 OC Y1 coincide with the middle plane
of WG and the plane xoy of HD planar model (see Fig. 3), respectively. Points O and OF are in the planes X1 OC Y1 and X2 OY2 ,
respectively. Y2 -axis and Y3 -axis are parallel to the plane Y1 OcZ1 and the plane Y2 OZ2 respectively. There are two angle
deviations between OC Z1 and OZ2 , which are respectively angle β 2 in the plane Y1 OC Z1 and angle β 3 in the plane X1 OC Z1 .
Similarly, angle α 2 and angle α 3 are the angular deviations between OZ2 and OF Z3 . As shown in Fig. 5, the eccentricity is
defined by length e1 , e2 and the angle β 1 , α 1 , and the deflection is defined by the four angle deviations (β 2 , β 3 , α 2 , α 3 ).
To investigate the effect of misalignment on pure kinematic error in the planar kinematic error model, the equivalent
WG profile curve is generated by projecting the WG profile in the plane X1 OC Y1 . The equivalent profile mainly depends on

6
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

the edge contour of the WG, which can be expressed in the coordinate system OF -X3 Y3 Z3 as

x23 y23 1
2
+ 2
= 1, z3 = ± b (16)
DW LW 2

where DW and LW are the semi-minor axis and semi-major axis of the elliptic WG, and b is the thickness of the WG.
Let P be a point on the contour curve of the WG and ri P = [xi , yi , zi , 1]T be the position vector of point P in the Oi (i = C,
F) coordinate system while rP = [x, y, z, 1]T be the position vector of point P in the O-system. The relationships among them
can be expressed in matrix form as

rCP = Ar P = ABrFP (17)


where
⎡− ⎤
(→ −

x 1 · x 2)
→ −
(− →
x 1 · y 2)
→ −
(− →
x 1 · z 2) a14
⎢( y 1 · →
→ −

x 2) ( y1·→
→ −

y 2) ( y1·→
→ −

z 2) a24 ⎥
A=⎣ − ⎦ (18a)
(→ −

z 1 · x 2) (−
→ − →
z 1 · y 2) (−
→ − →
z 1 · z 2) a34
0 0 0 1

→ − → −

where x 1 , y 1 and z 1 are unit vectors of axis X1 , Y1 and Z1 of coordinate systems OC -X1 Y1 Z1 , respectively. Similarly, the

→ − → −

x 2 , y 2 and z 2 are unit vectors of the axes of coordinate systems O-X2 Y2 Z2 . They can be expressed in coordinate system
OC -X1 Y1 Z1 as
⎧−
⎨→ x 1 = ( 1, 0, 0 )
T


y 1 = ( 0, 1, 0 )
T
(18b)
⎩−→
z 1 = ( 0, 0, 1 )
T

⎧    T
⎪ −→ 2 2 −0.5
⎪ sinβ2 tan β3
⎨ x 2 = 1 + ( cos β2 +tan β2 sin β2 ) + ( tan β3
) 1 , − sinβ2 tan β3
cos β2 +tan β2 sin β2
, − tan β3
1+tan β2
2
1+tan β2
2

−→ (18c)
⎪ y = ( 0 , cos β 2 , − sin β 2 )T

⎩− →
2
0.5 (tan β3 , tan β2 , 1 )
1 T
z2=
(tan2 β3 +tan2 β2 +1 )

a14 = e1 sin β1
a24 = e1 cos β1 (18d)
a34 = 0

The B in Eq. (17) can be expressed as


⎡− ⎤
(→ −

x 2 · x 3)
→ −
(− →
x 2 · y 3)
→ −
(− →
x 2 · z 3) b14

→ −
⎢( y 2 · →
x 3)

→ −
( y2·→ y 3)

→ −
( y2·→ z 3) b24 ⎥
B=⎣ − ⎦ (19a)
(→ −

z 2 · x 3) (−
→ − →
z 2 · y 3) (−
→ − →
z 2 · z 3) b34
0 0 0 1

→ − → −

where x 3 , y 3 and z 3 are unit vectors of axis X3 , Y3 and Z3 of coordinate systems OF -X3 Y3 Z3 , respectively. As previously

→ − → −

described, the x 2 , y 2 and z 2 are unit vectors of the axes of coordinate systems O-X2 Y2 Z2 . They can be expressed in
coordinate system O-X2 Y2 Z2 as
⎧−
⎨→x 2 = ( 1, 0, 0 )
T


y 2 = ( 0, 1, 0 )
T
(19b)
⎩−

z 2 = ( 0, 0, 1 )
T

⎧  −0.5  T

⎪−→ sinα2 tan α3 2 tan α3 2 α2 tan α3 tan α3
⎨ x 3 = 1 + ( cos α2 +tan α2 sin α2 ) + ( 1+tan2 α2 ) 1, − cos αsin
2 + tan α2 sin α2
, − 1+tan α2
2


→ (19c)
y 3 = (0, cos α2 , − sin α2 )
T


⎩→−
z3= 1
0.5 (tan α3 , tan α2 , 1 )
T
(tan2 α3 +tan2 α2 +1 )

b14 = e2 sin α1
b24 = e2 cos α1 (19d)
b34 = 0

Let Q be a point on the equivalent WG profile curve, and rQ = [x, y, z, 1]T be the position vector of point Q in the
OC -system. The equivalent WG profile curve lWG can be expressed in coordinate system OC -X1 Y1 Z1 as
 
Q
  x23 y23 1
r = max rCP , z =0 P∈ 2
+ 2
= 1 , z3 = ± b (20)
DW LW 2

7
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 6. Evaluation of the tangential composite deviation (source [31]).

3.2. Mathematical model of machining error for tooth

There is always a random error between theoretical tooth profile and the actual tooth profile considering the gear ma-
chining and machine tool accuracy. The machining error of gear is a comprehensive error, and can be divided into profile
error, helix error, and pitch error [27]. These errors have a great effect on the kinematic error of HD. In the planar kine-
matic model of HD, a comprehensive error indicator, tangential composite deviation, is used to analyze the contribution of
machining error to pure kinematic error. Fig. 6 demonstrates an experimental profile of the tangential composite deviation,
which can be broken down into two parts: tooth-to-tooth tangential composite deviation and tooth tangential composite
deviation. The tooth-to-tooth tangential composite deviation is short-wave components of tangential composite deviation,
and it is difficult to identify by a mathematical model. As shown in Fig. 6, it has the characteristics of small amplitude and
high frequency. Hence, its contribution to the error is not taken into account in this paper. The tooth tangential composite
deviation can be regarded as the long-wave components of tangential composite deviation, which can be normally expressed
as a sinusoidal function [30].
Fi = a sin(ωθ + ϕ ) (21)
where a, ω and ϕ represent the amplitude, frequency and phase angle of the tooth error, respectively.

4. Calculation of pure kinematic error for the harmonic drive

4.1. Nonuniform motion component of pure kinematic error

The instantaneous transmission ratio is not an eternally fixed constant for HD. It may be attributed to assembly and
machining imperfections on the three major parts and basic operating principle of drive. When the HD is in motion, the
nonuniform motion component acts as a periodic exciter and causes undesirable vibration effects. Next, the error is calcu-
lated by modeling and its mechanism is explained.

4.1.1. The wave generator misalignment owing to the assembly and machining imperfections
As shown in Fig. 7, in the plane X1 OC Y1 (see Fig. 5), the dynamic coordinate system xF OˊF yF is fixed on WG, the yF -axis
of the dynamic coordinate system is in coincidence with the major axis of WG, and point OˊF is the projection point of
point OF in the plane X1 OC Y1 . A static coordinate system xO OyO is fixed on the rotation center O. The yO -axis of the static
coordinate system is parallel to the Y1 -axis.
The equivalent WG profile curve lWG can be obtained by the Section 3.1. But for a clockwise input angle θ m , the
Eqs. (19c) and (19d) need to be modified as follows
⎧−→ H tan α3 T

⎨ x 3 = (I · sin(π − arcsin( I ) + θm ), I · cos(π − arcsin( I ) + θm ), − 1+tan2 α2 )
H H



y 3 = (cos α2 · sin θm , cos α2 · cos θm , − sin α2 )
T
(22a)

⎩−→ . 0.5 T
z 3 = ( (1 − J 2 ) sin(arcsin( J tan2α03.5 ) + θm ), (1 − J 2 ) cos(arcsin( J tan2α03.5 ) + θm ), J )
0 5
( 1− J ) ( 1− J )

b14 = e2 sin(α1 + θm )
b24 = e2 cos(α1 + θm ) (22b)
b34 = 0
where
⎧  −0.5

⎪H = 1 + ( cos αsin α2 tan α3 2
) + ( 1+tan α3 2
)

⎪ + tan α sin α 2
α
⎨  
2

2  0 . 5
2 2 tan 2

I = 1− H tan α3 (22c)

⎪ 1+tan2 α2


⎩J = 1
0.5
(tan2 α3 +tan2 α2 +1 )

8
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 7. Angles and frames of reference.

Fig. 8. Displacement increment along the FS reference line with misalignment in planar model.

Thus, for a clockwise input angle θ m , the corresponding radial displacement can be calculated by
  
δ (θm ) = max rQ  − LW , Q ∈ lWG (θm ) (23)
As shown in Fig. 8, if the WG has a misalignment, the meshing point P has a displacement along the CS tooth profile.
The angle ψ can be approximated as
δ tan α
ψ= (24)
RCS

9
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Table 1
Basic parameters of the HD.

Parameter Value Parameter Value

Number of FS teeth (ZFS ) 160 Number of CS teeth (ZCS ) 162


Module (m) 0.396 mm Pressure angle (α ) 12°
Major semi axis (LW ) 31.046 mm Minor semi axis (DW ) 30.318 mm
Width of WG (b) 9.00 mm Radius of FS inner wall (Rb ) 30.65 mm
FS addendum height 0.277 mm CS addendum height 0.257 mm
FS dedendum height 0.290 mm CS dedendum height 0.400 mm

Fig. 9. The kinematic error of HD with misalignment (ke = 1).

where δ is radial displacement, α is pressure angle. The equivalent displacement (arc length lAP ) of meshing point along
deformed FS reference circle can be expressed as
  2
 ψ
d ρ (θ )
lAP = ρ (θ )2 + dθ (25)
0 dθ

where ρ (θ ) denotes the polar representation of the deformed FS reference circle. According to the pure kinematic error
model in Section 2.2, the kinematic error owing to misalignment can be obtained
lAP ke
θ˜p1 = 2π   d ρ ( θ ) 2 (26)
 2π
0 ρ (θ )2 + dθ

Basic parameters of the HD (CSF-25–80) in this paper are shown in Table 1.


In order to study the influence of misalignment E1 (e1 = 0.01 mm, β 1 = 45°, β 2 = 0.2°, β 3 = 0.2°) and E2 (e2 =
0.005 mm, α 1 = 45°, α 2 = 0.1°, α 3 = 0.1°) on the kinematic error of HD, the kinematic error with E1 and E2 are calculated
separately, as is shown in Fig. 9. We note that there is no nonuniform motion component of kinematic error when only
misalignment E2 exists. This is because that the input axis coincides with the shaft axis of the CS, and the meshing state
of the CS teeth and the FS teeth is consistent for any input angle θ m . The misalignment E1 will cause an error mainly
occurring at twice the WG rotation (2ωb ). In order to research the comprehensive influence of the two misalignments on
the kinematic error, Fig. 9 shows the kinematic error of the HD with misalignment E1 and E2 . Compared to the results while
considering the misalignment E1 and misalignment E2 separately, the kinematic error adds a component mainly occurring
at once the WG rotation (ωb ). The peak-to-peak value of the kinematic error is 0.99 , which is bigger than the WG only with
misalignment E1 .

4.1.2. The tooth with machining error


For the most common configuration of the HD: fixed member CS; driven member FS; driving member WG, the tooth
tangential composite deviation for tooth of fixed CS can be described as a function of the WG input angle θ m as follows
 = a sin(θ + ϕ )
FiCS (27)
CS m CS

where aCS and ϕ CS represent the amplitude and phase angle of the CS tooth error, respectively. In this configuration, the WG
rotation corresponds to the input angle while the rotation of the FS along the opposite direction corresponds to a certain

10
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 10. The kinematic error of HD with tooth error (ke = 1).

output angle. The tooth tangential composite deviation for tooth of driven FS can be described as a function of the WG
input angle θ m as follows
Z 
FiF S = aF S sin θm + ϕ F S
CS
(28)
ZF S
where aFS and ϕ FS represent the amplitude and phase angle of the FS tooth error, respectively. The tangential composite
deviation is defined by the corresponding deviations in arc length of the reference circle between theoretical rotation angle
and nominal rotation angle. When WG misalignment does not exist, the two meshing areas are exactly opposite along two
directions of major axis of WG. According to the kinematic error model of Section 2.2, it was observed that only the tooth
with larger tangential deviation between the teeth with the maximum meshing depth plays a role in the pure kinematic
error. In practice, the pure kinematic error is affected by the error of CS and FS teeth simultaneously. The kinematic error
owing to teeth errors can be obtained
2π τ ke
θ˜p2 = (29a)
lFS
 Z  Z 
τ = max aCS sin(θm + ϕCS ) + aF S sin CS
θm + ϕF S , aCS sin(θm + ϕCS + π ) + aF S sin CS
θm + ϕ F S + π (29b)
ZF S ZF S
Considering the CS tooth error (aCS = 0.002 mm, ϕ CS = 0°) and FS tooth error (aFS = 0.002 mm, ϕ FS = 0°) separately,
Fig. 10 shows the kinematic error of HD with teeth errors. We can note that the kinematic error due to FS tooth error
has more high frequency, but their error curves are similar in shape. Synthesizing the CS tooth error and FS tooth error,
simulation results of the HD with teeth errors are shown in Fig. 10. And the error is increased by two components, one
mainly occurring at 2.0125 times the WG rotation (2.0125ωb ), the other mainly occurring at 0.0125 times the WG rotation
(0.0125ωb ). Relative position of FS and CS returns to initial state, when the FS rotates a single revolution. Hence it may
cause kinematic error component occurring at a frequency of ωb /N (0.0125ωb ).

11
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 11. Angles and frames of reference.

4.1.3. Synthesizing the effect of all errors on nonuniform motion component


Considering the comprehensive influence of WG misalignment and tooth error on nonuniform motion component, the
tooth with larger tangential deviation in two teeth with the largest meshing depth in their respective meshing zone plays a
role in the pure kinematic error. Therefore, kinematic error is not a linear superposition of errors caused by misalignment
and tooth error.
As shown in Fig. 11, the dynamic coordinate system XW OC YW is fixed on the ideal WG. For a clockwise input angle θ m ,
points H1 and H2 are the position with the largest meshing depth in the meshing zones 1 and 2, respectively, and their
coordinates in the coordinate system X1 OC Y1 can be obtained by the method of the Section 4.1.1. Let the coordinates of
points Hi in the dynamic coordinate systems XW OC YW and xF OˊF yF be (xWi , yWi ) and (xFi , yFi ), respectively. The transformation
matrix C from the system X1 OC Y1 to the system XW OC YW and the transformation matrix D from the system X1 OC Y1 to the
system xF OˊF yF can be expressed as
 
cos θm − sin θm 0
C= sin θm cos θm 0 (30a)
0 0 1
 
cos θm − sin θm −X1 (OF )
D= sin θm cos θm −Y1 (OF ) (30b)
0 0 1
where (X1 (OˊF ), Y1 (OˊF )) is the coordinate of point OˊF in system X1 OC Y1 , which can be obtained by the method proposed
in Section 3.1. Thus, the points Hi in the dynamic coordinate systems XW OC YW and xF OˊF yF can be obtained by coordinate
transformation.
From the geometry of Fig. 11, the equivalent displacement of meshing point along deformed FS reference circle in mesh-
ing zones 1 and 2 can be expressed separately as
Z 
τ1 = lAP1 + aCS sin(θm + arcsin(xW 1 /(xW 1 2 + yW 1 2 )0.5 ) + ϕCS ) + aF S sin CS
θm + arcsin(xF 1 /(xF 1 2 + yF 1 2 )0.5 ) + ϕF S
ZF S
(31a)

τ2 = lAP2 +aCS sin(θm + ϕCS + π − arcsin(xW 2 /(xW 2 2 + yW 2 2 )0.5 ))


Z 
+ aF S sin
CS
θm + ϕF S + π − arcsin(xF 2 /(xF 2 2 + yF 2 2 )0.5 ) (31b)
ZF S

12
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 12. The nonuniform motion component of pure kinematic error (ke = 1).

Based on the above discussions, the nonuniform motion component of pure kinematic error can be obtained

2 π τ1 k e / l F S τ1 > τ2
θme = (32)
2 π τ2 k e / l F S τ1 < τ2
Fig. 12 shows the nonuniform motion component of the HD with misalignment and tooth error (e1 = 0.01 mm,
β 1 = 45°, β 2 = 0.2°, β 3 = 0.2°, e2 = 0.005 mm, α 1 = 45°, α 2 = 0.1°, α 3 = 0.1°, aCS = aFS = 0.002 mm, ϕ CS = ϕ FS =
0°). The simulation results well agree with Tuttle [32] in that the profile of the kinematic error consists predominantly of
a fundamental harmonic occurring at a frequency of two cycles per WG rotation (2ωb ), and some other error components,
such as the error components with frequencies of ωb /N, ωb , 2(1 + 1/N)ωb and 2(1 + 1/2 N)ωb . The component of frequency
ωb /N reflects the comprehensive influence of the interaction between the CS tooth error and the FS tooth error on the
kinematic error. The value of the component of frequency ωb can reflect the severity of misalignment E2 , and the larger
is the value, the more serious is the misalignment E2 . It shows that the pure kinematic error curve will be observed beat
frequency phenomenon. This observation is due to the fact that the frequency (2(1 + 1/N)ωb , 2(1 + 1/2 N)ωb ) of the error
component caused by the FS tooth error is very close to that (2ωb ) of the fundamental error.
To research the effects of different misalignments, tooth errors, pressure angles, modules and widths of WG on nonuni-
form motion component, single variable method is adopted and the peak-to-peak values of the nonuniform motion error
are calculated (e1 = 0.01 mm, β 1 = 45°, β 2 = 0.2°, β 3 = 0.2°, e2 = 0.005 mm, α 1 = 45°, α 2 = 0.1°, α 3 = 0.1°, aCS = aFS =
0.002 mm, ϕ CS = ϕ FS = 0°). Looking at the calculations in Table 2, when eccentricity e2 is along the direction of the minor
axis of WG (α 1 = 90°), eccentricity e2 has the least effect on the kinematic error. Similarly, the deflection (α 3 ) along the di-
rection of the minor axis of WG has little effect on the kinematic error. However, the peak-to-peak value of error is obviously
affected by the deflection (α 2 ) along the major axis direction of WG. These can be attributed to the fact that the meshing
state of the CS teeth and the FS teeth mainly depends on the major axis of the WG. The error amplitude is more sensitive
to eccentricity e1 than that of eccentricity e2 . There is an eccentricity orientation to minimize the influence of eccentricity
e1 on the error, such as β 1 = 0° in the example. The deflection angles β 2 and β 3 have almost the same effect on the error
amplitude. In addition, we note that the error amplitude is very sensitive to the variation of manageable design parameters,

13
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Table 2
The nonuniform motion component of kinematic error (ke = 1).

Parameters Value Peak-to-peak value Parameters Value Peak-to-peak value


of nonuniform of nonuniform
error (arc min) error (arc min)

α 1 (degree) 0 1.36 e1 (mm) 0.008 1.28


45 1.33 0.01 1.33
90 1.25 0.012 1.38
α 2 (degree) 0.05 1.24 e2 (mm) 0.003 1.29
0.1 1.33 0.005 1.33
0.15 1.43 0.007 1.36
α 3 (degree) 0.05 1.33 α (degree) 10 1.17
0.1 1.33 12 1.33
0.15 1.33 14 1.50
β 1 (degree) 0 1.25 m (mm) 0.386 1.36
45 1.33 0.396 1.33
90 1.30 0.406 1.30
β 2 (degree) 0.15 1.31 b (mm) 8 1.26
0.2 1.33 9 1.33
0.25 1.41 10 1.40
β 3 (degree) 0.15 1.30 aCS 0.001 1.25
0.2 1.33 0.002 1.33
0.25 1.42 0.003 1.42
aFS 0.001 1.22
0.002 1.33
0.003 1.44

such as modulus, pressure angle and width of WG. It was observed that as the pressure angle and WG width decrease, the
error amplitude decreases. This is due to that the reduction of WG width weakens the influence of deflection on the error.
It is noted that the transmission accuracy can be improved by selecting large modulus. This conclusion actually agrees with
Nye and Kraml [8] in that the value of the error is inversely proportional to the pitch diameter. Summarily, the tooth error
has a more significant impact on the pure kinematic error than the WG misalignment. Specifically, the level of influence
on the pure kinematic error is in turn reduced by FS tooth error, CS tooth error, misalignment E1 and misalignment E2 . For
design parameters, the pressure angle and the width of WG are equally important on affecting the pure kinematic error, but
the influence of modulus on the kinematic error is small.

4.2. Lag component of pure kinematic error

The output lag characteristic is a disadvantage of HD particularly when it needs to reverse frequently. In this case, the
lag component plays an important role in the transmission accuracy. It occurs due to the clearance of contact pairs and the
torsional flexibility property of the drive. Under the ideal no-load condition, the torsional flexibility is not considered.

4.2.1. Kinematic error due to clearance


Clearance exists in various places such as gear teeth, bearings, and shaft assembly. The backlash between gear teeth
and the clearance of WG are considered in this study. As shown in Fig. 13, the common configuration of the WG includes
two types: automatic alignment and integral structures [33]. For the first structure, the automatic alignment mechanism is
presented in Fig. 13(b), which is designed according to the principle of the cross-slide mechanism. The clearance mainly
comes from the automatic alignment mechanism [33]. For the integral structure, the WG clearance is not considered. As
shown in Fig. 14, the geometric center OF of the FS coincides with the geometric center OC of the CS for HD with automatic
alignment WG.
Let eb be the clearance of WG (see Fig. 14). The dynamic coordinate system x1 OC y1 is fixed on WG. The y1 -axis of the
dynamic coordinate system is along the direction of the straight groove 1. Points Ox and Oy are the projection points of the
rotation center O on the x1 -axis and y1 -axis, respectively. The eccentricity of the rotation center O is defined by length e3
and angle β , and the angle β depends on input angle θ m . From the geometry of Fig. 14, for a clockwise input angle θ m , the
lag error due to the clearance of automatic alignment mechanism can be deduced

ϑt1 = ϑb1 + ϑb2 (33a)

eb 1
ϑb1 =  0 . 5  · N (33b)
 0 . 5 2
+ |e3 cos (β − θm )| + b1 /4
2 2
Rw − b1 /4
2

14
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 13. Structure of two wave generators: (a) Automatic alignment and integral structures, (b) Automatic alignment mechanism.

Fig. 14. Mechanism of kinematic error due to clearance of automatic alignment mechanism.

eb 1
ϑb2 =  0.5  · N (33c)
 0 . 5 2
+ |e3 sin (β − θm )| + b2 /4
2 2
Rw − b2 /4
2

where ϑb 1 and ϑb 2 are the lag error due to the clearance of straight groove 1 and 2, respectively.
Let et be the gear backlash. The time-varying gear backlash will be caused by misalignment and machining error of tooth.
But for the HD with automatic alignment WG, the gear backlash is only affected by the machining error of gear, and the lag
error can be expressed as

et − τ
ϑt2 = 2π (34)
lF S

where τ can be determined by Eq. (29b). For the HD with integral WG, the lag error can be obtained as follows

2π etl−τ1 τ1 > τ2
ϑt3 = FS
(35)
2π etl−τ2 τ1 < τ2
FS

where τ 1 and τ 2 are calculated by Eqs. (31a) and (31b) respectively.

15
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Table 3
Theoretical nonuniform motion component of error for different WGs.

WG type Peak-to-peak values of WG type Peak-to-peak values of


kinematic error (arc min) kinematic error (arc min)

Automatic alignment WG 1.0241 Major axis automatic alignment WG 1.0249


Integral WG 1.3290 Minor axis automatic alignment WG 1.3287

Fig. 15. Experimental torsional stiffness curves at different WGs.

5. Kinematic error and wave generator

The simulation results in Section 4.1.3 show that the kinematic accuracy can be improved by shortening the WG width.
In this section the influences of the WG structure and shape on the pure kinematic error will be investigated.

5.1. Wave generator structure

The pure kinematic error of HD is affected by WG structure. Here, two common configurations of WG (automatic align-
ment and integral structures) are investigated to analyze the pure kinematic error by the proposed method. The coefficients
used in the pure kinematic error simulation are displayed in Section 4.1.3. The theoretical results are shown in Table 3. It
can be seen that the peak-to-peak value of nonuniform motion component with automatic alignment WG is 1.0241 arcmin,
and the peak-to-peak value of nonuniform motion component with integral WG is 1.329 arcmin. This implies that nonuni-
form motion component of kinematic error can be optimized by designing an automatic alignment mechanism. On the other
hand, the automatic alignment mechanism will increase the lag component of error according to Section 4.2.1.
Moreover, the transmission performance of reducer is mainly evaluated by the transmission indices. For the HD, the
transmission indices mainly include kinematic error, transmission efficiency, and torsional stiffness. The results of torsional
stiffness of experiments are presented in Fig. 15 for the two types of WGs (Refer to Section 6 for test device and method).
As we observe in the figure, the stiffness for using integral WG is significantly larger than that for using the automatic
alignment WG. This may be due to the fact that the WG of the automatic alignment structure is less rigid than the integral
WG. On the other hand, the clearance of the automatic alignment mechanism weakens the torsional stiffness of the reducer.
The curves of efficiency with respect to input speed are shown in Fig. 16. We can see that the efficiency decreases with
the increase of the input speed. Similarly, the efficiency for using integral WG is larger than that for using the automatic
alignment WG. This may be attributed to the fact that the automatic alignment mechanism makes the transmission chain
longer and increases energy loss. In addition, kinematic error acts as an exciter and hence has undesirable vibration effects
especially for lower stiffness HD using the automatic alignment WG.
Based on the above analyses, we observed that the automatic alignment WG can eliminate the eccentricity to reduce the
nonuniform motion component of kinematic error. But it will be at the cost of weakening the stiffness and efficiency of the
reducer. In addition, it will increase the lag component of error, which will be obvious in forward-reversal rotation operating
conditions. For integral WG, it requires high machining and assembly accuracy, otherwise it will cause misalignment to
increase pure kinematic error and induce abnormal tooth wear.
This paper draws on the advantages of automatic alignment WG and integral WG. According to the analysis in
Section 4.1.3, the kinematic error is not sensitive to the misalignment in the minor axis direction of the WG. Therefore, we
try to propose a one-way automatic alignment WG to shorten the transmission chain and eliminate the one-way eccentric-
ity, and simulate the pure kinematic error with the proposed method. Table 3 shows the theoretical results of nonuniform
motion component for the two types one-way automatic alignment WGs: major axis and minor axis automatic alignment

16
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 16. Efficiency with respect to input speed.

Fig. 17. One-way automatic alignment mechanism.

(see Fig. 17, composed of input disc and output disc, without floating disc). It is noted that the major axis automatic align-
ment and the common automatic alignment WGs have almost the same improvement effect on the nonuniform motion
component of error. However, the minor axis automatic alignment WG has almost no improvement on the kinematic error
of the HD. It also implies that the major axis automatic alignment WG is feasible to improve kinematic error. In addition, it
shortens the transmission chain relative to the conventional automatic alignment WG, and reduces the weakening effects of
automatic alignment mechanism on lag component of error, torsional stiffness, and transmission efficiency.

5.2. Wave generator shape

As an effective design parameter, the selection of WG shape will affect the performance of the HD. In a general ar-
rangement, the external profile of the cam WG can be chosen as ellipse expressed in polar form by the following equation

(Rb + m )DW
ρ (θ ) =  (36)
(Rb + m )2 sin2 θ + DW 2 cos2 θ
where Rb is radius of the inner wall of the FS. The minor semi axis DW can be found by equating the FS length in deformed
and undeformed states.
In some designs, the WG is composed of two opposite rollers [9], and the WG shape is a Résal symmetric curve expressed
as
 
m 1 θ sin θ cos θ π π
ρ ( θ ) = Rb + − − − ≤θ ≤ (37)
1/π − π /8 π 4 4 2 2
In addition, the two-disk WG is also a common choice in design [26], as depicted in Fig. 18. The WG shape can be
expressed in polar form by the following equation
A1 cos θ −B1
ρ ( θ ) = Rb + 
A1 −B1
m
 0≤θ ≤γ
(38a)
ρ ( θ ) = Rb + m
A1 −B1
(
1 + sin2 γ )sinθ + ( π2 − θ )cosθ − 2sinγ − B1 γ < θ ≤ π /2
17
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 18. WG in the form of two-disk that apply a maximum radial inflection equal to the module m.

Table 4
Theoretical nonuniform motion component of error for different WG shapes.

Curve Single tooth meshing Maximum contact Minimum homogenization Minimum nonuniform
kinematic error ϑi angle (degree) coefficient ke (kB = 1) motion component of error
(arc min) θ me (arc min)
Ellipse 1.328 55.54 0.200 0.2656
Résal 1.329 54.99 0.201 0.2671
Two-disk (γ = 60°) 1.331 57.71 0.196 0.2609

where
π 4
! π  "
A1 = − γ − sin γ cos γ , B1 = cos γ − − γ sinγ (38b)
2 π 2
and γ is the angle of contact between the FS and the disk.
According to the error model proposed in this paper, on the one hand, the WG shape affects the kinematic error ϑi of
HD with single tooth meshing, which have been represented in Table 4. It is noted that the WG shape has little effect on
this part of error. On the other hand, the WG shape affects the homogenization coefficient that depends on the number of
meshing teeth. Under an external load, since the FS is essentially a compliant mechanism, neighboring teeth on the FS and
the CS may be forced to contact. The maximum contact angle can be determined by the intersection of the top circle of the
deformed FS and that of the CS, and the corresponding homogenization coefficient can be obtained by Eq. (13). As seen in
the Table 4, the homogenization coefficient corresponding to the two-disk WG is the smallest in the example. Summarily,
for these three different WG shapes, the corresponding kinematic errors are almost the same. This in turn implies that the
kinematic error is not sensitive to the change of WG shape.
Every WG shape has a different radius of curvature change with respect to the non-deformed FS, as shown in Fig. 19.
The change of curvature is directly related to the bending stress of the FS ring. As expected, the higher is the “contact ratio”
the higher is the bending stress. Hence, a fatigue limit in the “contact ratio” has to be set. The increase of “contact ratio”
may improve the load-carrying capacity, but as shown in Table 4, the improvement of kinematic error is very limited. In
addition, the simulation results with three different WG shapes show that the bending stress is more sensitive than the
“contact ratio” to the shape of the WG. Therefore, for the WG shape, its influence on bending stress should be paid more
attention to.

6. Model validation

6.1. Test apparatus

A specialized harmonic drive test device is designed to measure kinematic error, efficiency, and stiffness. As depicted in
Fig. 20, it is an electromechanical system that includes a servo drive motor, a load motor, rotary encoder, torque sensor, a HD
unit, and a digital control module. The concentricity is adjusted by an Easy-Laser shaft alignment instrument. The encoder
with resolution 0.00439° is selected at the input end. A higher-precision encoder (resolution = 0.0 0 056°) is used for the

18
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 19. Radius of curvature change vs. θ .

Fig. 20. View of the harmonic drive test apparatus.

load position, considering that the rotation speed of the output end is obviously slower than the input speed. In order to
measure the stiffness of reducer and transmission efficiency, the torque sensors are installed at the input and output sides
respectively.
For the measurement process of pure kinematic error with no load, the coupling 3 in the test bench is disengaged. In
order to suppress the influence of vibration excitation on pure kinematic error, the test needs to be carried out at a low
speed. According to the research of Ghorbel et al. [14], rotating the WG at a speed of less than 10 r/min is considered
enough to avoid the vibration excitation due to flexibility. In this study, when the input speed is stable at 5 r/min, the input
and output angle values are synchronously collected in real time, and the pure kinematic error is calculated by Eq. (8). To
determine the stiffness of HD experimentally, first the input shaft of the drive is locked. Next, the load torque is controlled
to vary in a liner manner, and the resulting angular displacement for the load position is measured by the rotary encoder
and is recorded.

6.2. Verification of suitability

Fig. 21 shows experimental profile of the pure kinematic error under input angle forward and reverse. It was observed
that a phenomenon similar to up and down translation appears on the pure kinematic error curve before and after changing
the rotation direction of input shaft. This phenomenon is interpreted by the lag component of the pure kinematic error.
Moreover, the error of one-way motion is seen as the nonuniform motion component of the kinematic error. Based on the
above analysis of the test results, in this study, it is reasonable to decompose the pure kinematic error into two components.
To prove the suitability of the proposed approach, a HD (CSF-25–80) with integral WG is used to test nonuniform mo-
tion component with no load on the specialized harmonic drive test device, the kinematic error of HD is calculated by the

19
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Fig. 21. Waveform of pure kinematic error under input angle forward and reverse.

Fig. 22. Simulation results of the kinematic error.

Table 5
The coefficients used in the kinematic error simulation.

Name Input value Name Input value Name Input value

e1 (mm) 0.01 e2 (mm) 0.0015 aCS (mm) 0.0015


β 1 (°) 70 α 1 (°) 80 aFS (mm) 0.0025
β 2 (°) 0.15 α 2 (°) 0.02 ϕ CS (°) 40
β 3 (°) 0.2 α 3 (°) 0.1 ϕ FS (°) 50

proposed approach (ke = 1), and then the numerical solution of the approach is compared with the testing result. The per-
tinent parameters of HD are shown in Table 1. The coefficients used in the kinematic error simulation cannot be measured,
which are determined to match experimental results and shown in Table 5. Fig. 22 shows the estimated and measured
kinematic error curves. The estimated kinematic error curves, which replicate the shape of HD typical pure kinematic error
curves, are shown to coincide with those obtained experimentally, and the suitability of the proposed approach is proved.
The mismatch between these two shapes stems from the fact that the tooth tangential composite deviation is assumed
a sinusoidal function, whereas the deviation cannot be expressed as a perfect sinusoidal function, and the tooth-to-tooth
tangential composite deviation of the small amplitude and high frequency is not taken into account.
According to the presented kinematic error model in this study, the frequencies of pure kinematic error can be obtained
as shown in Table 6. The calculated frequencies of pure kinematic error are shown to closely coincide with those found
experimentally. This indicates that the mechanism explanation of pure kinematic error is reasonable. Moreover, the table
shows that the linear superposition of each error cannot well reflect the combined effects on kinematic error of the inter-
action of the individual errors.

20
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

Table 6
The main frequencies of nonuniform motion component.

Sources of error Categories of error Separate impact Synthetical kinematic error frequency

Misalignment Misalignment E1 2ω b ωb /N, ωb , 2ωb , 2(1 + 1/N)ωb ,


Misalignment E2 No 2(1 + 1/2 N)ω b
Tooth error FS teeth 2(1 + 1/N)ω b
CS teeth 2ω b

7. Conclusions

Herein, synthesizing the machining errors and assembly errors, the calculation model of the pure kinematic error of the
HD is proposed, and the error is decomposed into two parts: nonuniform motion component and lag component. A special-
ized harmonic drive test device is designed to measure pure kinematic error, and the reasonability of proposed approach is
demonstrated experimentally. The main conclusions are shown in the following:
(1) Due to the existence of random machining and assembly errors, the characteristics of pure kinematic error are fluc-
tuated perceptibly. The influence of machining and assembly errors along the major axis of WG on pure kinematic
error variations of HD is worse than other axis cases.
(2) The tooth error has a more significant impact on the pure kinematic error than the WG misalignment. The pressure
angle and the width of WG are equally important on affecting the pure kinematic error, but the influence of modulus
on the kinematic error is small.
(3) Because the frequency of the error component caused by the FS tooth error is very close to that of the fundamental
error, pure kinematic error is observed with obvious beat frequency characteristic.
(4) A reasonable reduction in the pressure angle and WG width can reduce the peak-to-peak value of the pure kinematic
error. In addition, the selection of large modulus can also improve transmission accuracy.
(5) The pure kinematic error is not sensitive to the change of WG shape, but the WG shape is the main factor of the FS
bending stress.
(6) The traditional automatic alignment mechanism can reduce the amplitude of the nonuniform component of the kine-
matic error, but at the same time, it worsens the stiffness, efficiency and error lag component of the HD. It is found
that the major axis automatic alignment mechanism has the same improvement effect on the error nonuniform com-
ponent, while reducing the negative impact of the automatic alignment mechanism.

Declaration of Competing Interest

We would like to submit the enclosed manuscript entitled “Modeling and analysis of pure kinematic error in harmonic
drive”, which we wish to be considered for publication in “Mechanism and Machine Theory”. All co-authors have seen and
agree with the contents of the manuscript and there is no conflict of interest. We certify that the submission is original
work and is not under review at any other publication.

Acknowledgments

The present study is partially supported by National Key R&D Program of China (2018YFB1304800), and the Basic science
and frontier technology research of Chongqing (cstc2018jcyjAX0363).

References

[1] X. Chen, Y. Liu, J. Xing, S. Lin, W. Xu, The parametric design of double-circular-arc tooth profile and its influence on the functional backlash of harmonic
drive, Mech. Mach. Theory 73 (2014) 1–24.
[2] E. Brassitos, N. Jalili, Design and development of a compact high-torque robotic actuator for space mechanisms, ASME J. Mech. Robot. 9 (6) (2017)
061002.
[3] Z. Shi, Y. Li, G. Liu, Adaptive torque estimation of robot joint with harmonic drive transmission, Mech. Syst. Signal Process. 96 (2017) 1–15.
[4] A.F. Emel’yanov, Calculation of the kinematic error of a harmonic gear transmission taking into account the compliance of elements, Sov. Eng. Res. 3
(7) (1983) 7–10.
[5] T. Tuttle, W. Seering, Kinematic error, compliance, and friction in a harmonic drive gear transmission, The 1993 ASME Design Technical Conferences-
19th Design Automation Conference, pp. 319–324.
[6] L. Hsia, The analysis and design of harmonic gear drives, Proceedings of the 1988 IEEE International Conference on Systems, Man, and Cybernetics,
pp. 616–619.
[7] R. Ramson, Positional error analysis of harmonic drive gearing, Master’s Thesis, Clemson University, 1988.
[8] T. Nye, R. Kraml, Harmonic drive gear error: characterization and compensation for precision pointing and tracking, in: Proceedings of the 25th
Aerospace Mechanics Symposium, 1991, pp. 237–252.
[9] F. Gravagno, V.H. Mucino, E. Pennestrì, Influence of wave generator profile on the pure kinematic error and centrodes of harmonic drive, Mech. Mach.
Theory 104 (2016) 100–117.
[10] H. Dong, T. Chen, D. Wang, B. Dong, Kinematic model of harmonic drive in robot joints with input eccentricity error, J. Dyn. Control Syst. (2019)
134–140.
[11] M. Yamamoto, M. Iwasaki, H. Hirai, Y. Okitsu, K. Sasaki, T. Yajima, Modeling and compensation for angular transmission error in harmonic drive
gearings, IEEJ Trans. Electr. Electron. 4 (2) (2009) 158–165.

21
H. Jia, J. Li, G. Xiang et al. Mechanism and Machine Theory 155 (2021) 104122

[12] T.D. Tuttle, W.P. Seering, A nonlinear model of a harmonic drive gear transmission, IEEE Trans. Robot. Autom. 12 (3) (1996) 368–374.
[13] H. Zhang, S. Ahmad, G. Liu, Torque estimation for robotic joint with harmonic drive transmission based on position measurements, IEEE Trans. Robot.
31 (2) (2015) 322–330.
[14] F.H. Ghorbel, P.S. Gandhi, F. Alpeter, On the kinematic error in harmonic drive gears, J. Mech. Des. 123 (2001) 90–97.
[15] D. León, N. Arzola, A. Tovar, Statistical analysis of the influence of tooth geometry in the performance of a harmonic drive, J. Braz. Soc. Mech. Sci. 37
(2) (2015) 723–735.
[16] Y. Liu, C. Tan, Y. Zhao, H. Liu, Y. Liu, Nonlinear attributes modeling and analysis of harmonic drive manipulator joint, in: International Conference on
Control, IEEE, 2017, pp. 256–264.
[17] T. Tjahjowidodo, F. Al-Bender, H. Van Brussel, Theoretical modelling and experimental identification of nonlinear torsional behaviour in harmonic
drives, Mechatronics 23 (5) (2013) 497–504.
[18] R. Dhaouadi, F.H. Ghorbel, P.S. Gandhi, A new dynamic model of hysteresis in harmonic drives, IEEE Trans. Ind. Electron. 50 (6) (2003) 1165–1171.
[19] H. Zhang, S. Ahmad, G. Liu, Modeling of torsional compliance and hysteresis behaviors in harmonic drives, IEEE-ASME Trans. Mech. 20 (1) (2014)
178–185.
[20] P.S. Gandhi, F.H. Ghorbel, J. Dabney, Modeling, identification, and compensation of friction in harmonic drives, in: IEEE Conference on Decision and
Control, 2002, pp. 160–166.
[21] P. Curt, R.J. Thomas, S. Deming, A high-fidelity harmonic drive model, J. Dyn. Syst-Trans. ASME 134 (1) (2012) 011002.
[22] P.S. Gandhi, F.H. Ghorbel, Closed-loop compensation of kinematic error in harmonic drives for precision control applications, IEEE Trans. Control Syst.
Technol. 10 (6) (2002) 759–768.
[23] R. Bikash, Design aspects of harmonic drive gear and performance improvement of its by problems identification: a review, in: American Institute of
Physics Conference Series, American Institute of Physics Conference Series, 2018.
[24] H. Dong, K. Ting, D. Wang, Kinematic fundamentals of planar harmonic drives, J. Mech. Design 133 (1) (2011) 011007.
[25] H. Dong, D. Wang, K. Ting, Kinematic effect of the compliant cup in harmonic drives, ASME J. Mech. Design 133 (5) (2011) 051004.
[26] X. Chen, Y. Liu, J. Xing, S. Lin, M. Ma, A novel method based on mechanical analysis for the stretch of the neutral line of the flexspline cup of a
harmonic drive, Mech. Mach. Theory 76 (2014) 1–19.
[27] T. Lin, Z. He, Analytical method for coupled transmission error of helical gear system with machining errors, assembly errors and tooth modifications,
Mech. Syst. Signal Pr. 91 (2017) 167–182.
[28] J. Li, Failure mechanism theory and accelerated life testing method research for space lubrication harmonic drive, Chongqing University. 2012. (PhD
thesis)
[29] M.N. Ivanov, The Harmonic Drive, Defense Industry Press, Beijing, 1987.
[30] M. Pueo, J. Santolaria, R. Acero, A. Gracia, A review of tangential composite and radial composite gear inspection, Precis. Eng. 50 (2017) 522–537.
[31] ISO 1328-1:2013 Cylindrical gears – ISO system of accuracy – Part 1: definitions and Allowable Values of Deviations Relevant to Corresponding Flanks
of Gear Teeth, 2013.
[32] T.D. Tuttle, Understanding and modeling the behavior or a harmonic drive gear transmission, Artificial Intelligence Laboratory, Massachusetts Institute
of Technology, 1992. (PhD thesis)
[33] Y. Shen, Q. Ye, Theory and Design of Harmonic Drive, China Machine Press, Beijing, 1985.

22

You might also like