1 s2.0 S0022460X23001177 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Journal of Sound and Vibration 553 (2023) 117668

Contents lists available at ScienceDirect

Journal of Sound and Vibration


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

Decoupling design and control of a spaceborne ultra–stable


platform for vibration isolation and precise steering
Chao Liang a, Weipeng Li a, *, Hai Huang a, Zhuokun Wu b
a
School of Astronautics, Beihang University, 37 Xueyuan Road, Beijing 100191, China
b
53rd Research Institute of CETC, Tianjin 300308, China

A R T I C L E I N F O A B S T R A C T

Keywords: This study concerns the problems of system coupling and large payload mass ratio and designs a
Decoupling design spaceborne ultra–stable platform (SUSP) to isolate the payload from the satellite vibration and
Vibration isolation accomplish high–precision payload steering. First, the dynamic equilibrium of the SUSP and vi­
Steering control
bration transfer functions are determined. The decoupling condition is then derived to make the
Stewart platform
Spacecraft dynamic
coupled vibration transfer functions equal to zero, and the SUSP multi–input multi–output
(MIMO) system is decoupled into six single–input single–output (SISO) subsystems. Consequently,
the natural frequency matrix can be easily designed to guarantee the vibration isolation perfor­
mance and obtain identical feedback gain values. Subsequently, a steering controller based on a
nominal system simplified by decoupled dynamic equilibrium was designed. The impact of a large
payload mass ratio was estimated and eliminated by the improved disturbance observer (IDO) in
the steering controller. After rigid–flexible coupling dynamic modeling and analysis, two simu­
lation stages were implemented. The results of the isolation stage exhibit a –40 dB/dec roll–off of
the vibration disturbance and a coupled response reduction of approximately –20 dB. The payload
tracking error of the isolation–steering integrated stage is reduced to 0.13′′ (0.1◦ circular trace),
and the isolation performance is further improved.

1. Introduction

With the development of spacecraft applications, various spaceborne payloads, such as astronomical telescopes, satellite antennas,
space interferometers and laser communication devices, have higher requirements for the accuracy and stability of payload pointing
[1]. However, pointing errors can be increased by micro–vibration disturbances, resulting from on–board devices such as reaction
wheels assembly (RWA), control momentum gyroscopes (CMGs), solar array driving assemblies (SADAs), flexible appendage, etc. [2],
with low magnitude and wide frequency bands from below 1 Hz to above 1 kHz [3]. Small errors also lead to target loss or imaging
jitter because of the long distance [4], which sets forth the goal of combining vibration isolation and steering.
Passive isolation is the main choice for resonance attenuation and high–frequency vibration isolation (30 Hz to > 1 kHz) without
any external energy [5]. However, a passive isolator is not effective for low–frequency vibrations (< 1 Hz); thus, an active isolator with
control–force inputs is suitable for improving the system performance in the low–frequency band range. This combination of active and
passive isolators creates a hybrid isolator [6]. Increasing attention has been paid to hybrid isolation based on the Stewart platform (or
hexapod) [7] owing to its high structural stability, large load capacity, and small error accumulation [8]. Consequently, numerous

* Corresponding author.
E-mail address: liweipeng@buaa.edu.cn (W. Li).

https://doi.org/10.1016/j.jsv.2023.117668
Received 14 September 2022; Received in revised form 1 January 2023; Accepted 8 March 2023
Available online 12 March 2023
0022-460X/© 2023 Elsevier Ltd. All rights reserved.
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

studies have applied the Stewart platform for vibration isolation and precise steering for spaceborne payloads [9–20]. However, these
studies did not focus on the impacts of system coupling and payload mass ratio, which can affect the system performance.
System coupling is a core problem in the Stewart platform. A previous study [21] demonstrated that a coupled system is less
effective than a decoupled system in terms of error compensation, robustness and stability. For a normal multiple–input multi­
ple–output (MIMO) system, both the input and output include six degrees–of–freedom (6 DOFs, including TX, TY, TZ, RX, RY, and RZ,
which denote translation along the X–, Y–, and Z–axes and rotation around the X–, Y–, and Z–axes, respectively) components. Coupling
induces multi–DOF coupled responses despite a single–DOF vibration input, which causes difficulties in designing the controller and
reduces the vibration control performance and steering accuracy [22]. For instance, when the satellite vibration in the TY DOF, xBTY , is
attached to the system, coupling will cause not only the payload vibration response in the TY DOF, xPTY , because the RX–DOF response,
xPRX , is also aroused as shown in Fig. 1, which can be termed TY–RX coupling. A “Cubic” architecture [9] was adopted to minimize the
coupling between struts, but the system coupling remained unresolved. The controller and decoupling designs are two different ap­
proaches for eliminating coupling effects. The controller design usually reduces the coupled vibration responses by developing specific
a control strategy for a coupled system. Yang et al. [23] added position sensors in struts to compensate for the coupled responses, which
increases the system complexity and power requirements. The decoupling design obtains a system without coupled responses through
numerical or analytical methods, which is termed decoupled system. Although a numerical decoupling design [24–26] has wide
applicability owing to various constraint of architecture design, it is still less efficient than analytical methods. McInroy et al. [21,27,
28] investigated analytic methods to design a decoupled architecture by first constructing a Jacobian matrix, while ignoring the
practical mounting space; however, the lack of physical intuition makes it difficult to implement the designed architecture in actual
situations [29]. In contrast to the aforementioned research, this study (1) presents the decoupling condition and reveals the physical
intuition of the decoupling concept, (2) provides an analytical rather than numerical method to design a decoupled system, avoiding
large amount of calculation, and (3) directly obtains the decoupling condition from the architecture parameters, which allows
considering the physical design constraints first and thereby guarantees the realizability of the designed decoupled system.
Another problem is that the payload mass ratio increases as payload function requirements become higher. The most well–known
space telescopes have a large payload mass ratio, i.e., the Spitzer Space Telescope has a payload mass of 851.5 kg and total mass of 950
kg [30], and the James Webb Space Telescope has a launch mass of 6500 kg and payload mass of 6330 kg [31]. A satellite and payload
can be regarded as a floating two–body system. When the payload is driven to track a moving target, a counterforce, which is usually
negligible when the payload mass ratio is small, acts on the satellite. However, when the payload mass ratio is large, the counterforce
becomes sufficiently large to cause a non–negligible satellite jitter in the low–frequency band range. Due to the connection between the
payload and satellite, the jitter will be further transferred to the payload and will take long time to decay [32]. Li et al. [33] developed a
pointing and vibration control platform for a 1000 kg optical payload and 1500 kg satellite with a Proportion Integration and Dif­
ferentiation (PID) controller, but it is applicable to static pointing rather than tracking. Kong et al. [34] studied a spacecraft with same
mass (600 kg) and proposed a dual–stage actuation that results in double energy consumption. Another study [35] used a multi–DOF
manipulator and non–contact actuators to separate the payload from the satellite. However, the manipulator enhances the system
uncertainties such as structural–flexibility and transmission errors. Compared with the aforementioned research, this study recognizes
the jitter caused by the large payload mass ratio as a state variable and designs a steering controller with improved disturbance
observer (IDO) to eliminate the jitter based on the nominal system simplified by the decoupling design, which improves the tracking
accuracy and avoids increasing the system complexity.
This study aims to provide a design method for decoupling the MIMO system and a controller for eliminating the impact of the large
payload mass ratio and use the idea to design a spaceborne ultra–stable platform (SUSP) to achieve an in–orbit vibration isolation and

Fig. 1. TY–RX coupling.

2
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

accurate payload steering. The SUSP is based on the Stewart platform architecture and is mounted between the satellite and payload, as
shown in Fig. 2(a). As shown in Fig. 2(b), the SUSP is composed of a mobile plate (MP, fixed to the payload), a base plate (BP, fixed to
the satellite), and six parallel variable–length struts (including actuators as well as upper and lower joints). The main contributions of
this study include the following: (1) an analytical method is provided to decouple the SUSP MIMO system into six single–input sin­
gle–output (SISO) subsystems as shown in Fig. 3, and reduce the coupled vibration responses; (2) the modeling is simplified by the
decoupling design and verified by the finite element analysis (FEA), and thereby the distributed steering controller can be developed
for the separate steering of RX and RY DOFs without affecting the remaining DOFs; (3) the disturbance observer is improved to es­
timate and eliminate the jitter caused by the large payload mass ratio and expand the vibration isolation bandwidth of the SUSP.
The remainder of this paper is organized as follows: Section 2 discusses the derivation of the decoupling condition and design of the
decoupled SUSP system. Section 3 discusses the development of SUSP isolation and steering controllers. In Section 4, the rigid–flexible
coupling dynamic model of the SUSP is established. Sections 5 discusses simulations conducted to verify the effectiveness of the
decoupling design and controller, as well as the vibration isolation and steering functions of the SUSP. Finally, Sections 6 presents the
conclusion.

2. Decoupling design of an SUSP

The decoupling condition used to decouple the SUSP MIMO system is derived in this section. Subsequently, the results of the
decoupled SUSP design are presented, and a 6–DOF vibration transmissibility analysis is conducted to demonstrate the effectiveness of
the decoupling design.

2.1. Decoupling condition

To obtain the decoupling condition, a numerical model of the SUSP should be established first. As the elongations of the struts are
very small owing to the in–orbit micro–vibration disturbances and small steering angle, the positions of the struts (joints and actuators)
can be assumed to be unchanged. Thus, it is reasonable to establish the SUSP numerical model in the initial positions of the struts,
which can be identified by the architecture parameters listed below, as shown in Fig. 2(b):

(1) Base plate radius: rb .


(2) Mobile plate radius: rp .
(3) Lower joint distribution angle: αb .
(4) Upper joint distribution angle: αp .
(5) Architecture height: h.

The payload frame {P} of the SUSP is established at a point fixed to the MP, which called the manipulation center (CP ), and the base
frame {B} is established with the origin at the center of the BP. Both the X–axes of {P} and {B} are parallel to the angle bisector (dashed
line) of αp as Fig. 2(b). Because CP is the rotation center for the SUSP payload steering task, the center of the payload mass (with the
MP) is designed to coincide with CP to decrease the moment of inertia. Therefore, CP should be carefully positioned to acquire the
desired vibration responses and improve the control performance.
The 6–DOF vibration disturbances, XB = (xBTX , xBTY , xBTZ , xBRX , xBRY , xBRZ )T , from the satellite can be transferred to the payload

Fig. 2. (a) SUSP mounted between the satellite and payload, and (b) SUSP configuration.

3
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

Fig. 3. Decoupling design target of an SUSP system.

(CP ) by causing telescopic motion of the six struts. The transmission between the strut motion and CP can be expressed as the Jacobian
matrix JP :
⎡ ⎤
sT1 (p1 × s1 )T
⎢ ⎥
JP = ⎢
⎣⋮ ⋮ ⎥,
⎦ (1)
sT6 (p6 × s6 )T

⇀ ⇀ ̅̅→
where si = bi pi /|bi pi | denotes a unit direction vector along the ith strut, and pi = CP pi = (pxi , pyi , pzi )T denotes the vector pointing to
the ith upper joint pi with respect to CP ; i = 1, 2, ⋯, 6.
Let XP = (xPTX , xPTY , xPTZ , xPRX , xPRY , xPRZ )T be the CP 6–DOF motion. The dynamic equilibrium of the payload is established as

MP ẌP = − K(XP − XB ) − C(ẊP − ẊB ) + FP , (2)

where, MP is the mass matrix of the payload (with the MP), which is regarded as a rigid body; K and C are the stiffness and passive
damping matrix of the SUSP, respectively; and FP is the 6–DOF control force attached to the payload, which is converted from the
actuation force of each strut. For the SUSP MIMO system, XP is the control output, and XB and FP are the disturbance and control inputs
respectively.
The stiffness matrix K is obtained as follows:

K = kJTP JP . (3)

where k denotes the axial stiffness of the struts. The passive damping matrix C was neglected because of the small structural damping
[36]. Eq. (2) can be written as:

MP ẌP + kJTP JP XP = kJTP JP XB + FP . (4)

FP consists of two parts: the vibration–isolation force FPiso and the steering force FPstr .

FP = FPiso + FPstr = JTP ⋅ Faiso + JTP ⋅ Fastr , (5)

where, Faiso and Fastr are the actuation forces for isolation and steering, respectively, and Faiso is calculated using the force–feedback
control algorithm [9]:
∫t
Faiso = − g Fl dt, (6)
0

where Fl is the force feedback measured by the force sensors installed along the struts and g is the force feedback gain. In space
environment, the force along the struts is the only force applied to the payload. Thus, FPiso is related to the absolute velocity of payload
as Fl is proportional to the payload’s acceleration:
∫t
( T )
FPiso = − g JP Fl dt = − g ⋅ MP ẊP . (7)
0

Combining Eqs. (4), (5) and (7), the dynamic equilibrium can be derived as

MP ẌP + gMP ẊP + kJTP JP XP = kJTP JP XB + FPstr . (8)

4
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

In this study, assume that the payload mass matrix MP is already designed to be diagonal. According to Eq. (8), if the matrix GJ =
JTP JP becomes diagonal, the single–DOF vibration–disturbance input or the single–DOF steering–force input can only cause the payload
response output in the same DOF, which implies that the MIMO system is decoupled into six SISO subsystems.
According to the rotation symmetry of the Stewart platform, it is reasonable to place CP on the Z–axis of {B} to maintain symmetry.
Therefore, the CP location is the height along the Z–axis of {B}, zc , and the first strut component in JP is
⎧ T
⎨ s1 = (sinθcosφ, sinθsinφ, cosθ)

(
1 1
)T , (9)

⎩ p1 = rp cos αp , rp sin αp , pz1
2 2

where, θ is the angle between s1 and the positive Z–axis, and φ is the angle between the projection of s1 on the XY plane and the positive
X–axis, which can be calculated using the architecture parameters and CP location. Using mirror symmetry, the second strut
component can be expressed as follows:
⎧ T
⎨ s2 = (sinθcosφ, − sinθsinφ, cosθ)

(
1 1
)T . (10)

⎩ p2 = rp cos αp , − rp sin αp , pz1
2 2

The other strut components can be obtained by rotation:


s3 = Rz,23π s1 , p3 = Rz,23π p1 ,
s4 = Rz,23π s2 , p4 = Rz,23π p2 ,
(11)
s5 = Rz,43π s1 , p5 = Rz,43π p1 ,
s6 = Rz,43π s2 , p6 = Rz,43π p2 ,

where, the rotation matrix is


⎡ ⎤
cosϑ − sinϑ
Rz,ϑ = ⎣ sinϑ cosϑ ⎦. (12)
1

By combining Eq. (9)–(12), the general form of GJ can be written as


⎡ ⎤
ηTX ηTXRY
⎢ ηTY ηTYRX ⎥
⎢ ⎥
⎢ ηTZ ⎥
GJ = ⎢

⎥,
⎥ (13)
⎢ ηTYRX ηRX ⎥
⎣η ηRY ⎦
TXRY
ηRZ

where, ηTX , ηTY , ηTZ , ηRX , ηRY , and ηRZ characterize the transfer coefficients of the SUSP in each DOF, and ηTYRX = − ηTXRY is the
relationship between coupled transfer coefficients in the TY and RX DOF, and those in the TX and RY DOF, respectively. ηTYRX is
derived as
[ ( ) ]
1
ηTYRX = 3sinθ rp cosθcos αp − φ − pz1 sinθ . (14)
2

When ηTYRX = 0, GJ become diagonal, which is only fulfilled when one of the following equivalents is ensured:

(1) sinθ = 0

Then, θ = 0, which is a singular pose of SUSP, where all struts are parallel to the Z–axis. Therefore, sinθ = 0 cannot be satisfied.
( )
(2) sinθ ∕= 0 and rp cosθcos 12αp − φ − pz1 sinθ = 0Thus, the decoupling condition is given by
( )
1
p∗z1 = rp cotθcos αp − φ . (15)
2
The optimal location of CP along the Z–axis of {B} can be obtained as follows:
z∗c = h − p∗z1 . (16)

The decoupling condition presented can be interpreted through physical intuition. Fig. 4 shows the brief architecture with only
strut 1 and 2, where the blue and black circles denote the MP and BP, respectively, the little red circles denote the joints and the red

5
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

arrows denote the struts. As Fig. 4(a), the point t1 moves along the strut 1, CP moves along the Z–axis of {B}, and the line CP t1 maintains
horizontal. When CP t1 ⊥p1 t1 , obtain point p∗1 by projecting p1 onto the XY plane of {P}. Thus, CP t1 ⊥p∗1 t1 .
In triangle ▵t1 p∗1 CP , αp /2 − φ − ∠t1 p∗1 CP = − π and |CP p∗1 | = rp , and then obtain the length of line p∗1 t1 is |p∗1 t1 | = |rp cos(αp /2 − φ)|,
as Fig. 4(b) shows. Considering that the length of line p∗1 p1 is |p∗1 p1 | = |p∗1 t1 | ⋅ cotθ, CP is at present on the optimal location as Eqs. (15)
and (16).
From the rotation symmetry, the struts are lying on a circular hyperboloid. Draw the yellow dashed circle with the center of CP and
the radius of rt = |CP t1 |, which is the throat of the hyperboloid. The decoupling condition given by Eqs. (15) and (16) reveals that the
optimal location of CP is at the center of the throat of the hyperboloid, which can be used to further simplify the decoupling design
process.
If CP is not strictly on the Z–axis, GJ becomes complicated and cannot be diagonal. The modified GJ formulas are provided in
Appendix A, which can be used to evaluate the degree of coupling caused by the deviation from the Z–axis. When the deviation is small,
the decoupling condition can be approximately satisfied.

2.2. Decoupling design

Using the decoupling condition given in Eqs. (15) and (16), a decoupled SUSP considering the actual payload interfaces and en­
velope requirements is designed. The SUSP architecture is displayed in Fig. 5, where the red lines denote the six struts with upper and
lower joints located at the two ends, the blue and black circles denote the MP and BP, respectively, and the black point denotes CP .
According to the discussion in Section 2.1, the throat tangents to struts 1 and 2 at t1 and t2 , respectively. Make both of the struts
perpendicular to the X–axis as Fig. 5(a), and then t1 and t2 coincide. Thus, the optimal location of CP can be easily found, which
coincides with the height of the intersection point of the two struts, as the green dashed lines show in Fig. 5(b). Then, the optimal
location of CP can be controlled by adjusting θ. The main parameters of the SUSP are listed in Table 1. In a real situation, the payload
mass includes the payload and MP. k is easily designed according to the natural frequency demand and payload mass matrix of the
SUSP, as the SUSP is now decoupled into six single–DOF SISO subsystems.

2.3. Transmissibility analysis

To demonstrate the effectiveness of the decoupling design in improving the vibration transfer characteristics and simplifying the
dynamic model, the vibration transfer functions between XB and XP were deduced as shown below. Eq. (8) can be rewritten in the
general form as

Fig. 4. Brief architecture with only strut 1 and 2, (a) 3–D diagram and (b) top view.

6
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

Fig. 5. Architecture design of SUSP, (a) top view, (b) side view and (c) 3–D diagram.

Table 1
Main parameters of SUSP.
Items Values

Base plate radius rb = 561.153 mm


Mobile plate radius rp = 534.248 mm
Lower joint distribution angle αb = 36.452◦
Upper joint distribution angle αp = 7.833◦
Architecture height h = 139.017 mm
CP location z∗c = 175.508 mm
Mass of payload m = 1005.7 kg
Inertia of payload IXX = 134.6 kg × m2,
IYY = 134.3 kg × m2,
IZZ = 143.3 kg × m2
Strut axial stiffness k = 27027 N × m − 1

⎡ ⎤
⎡ ⎤ ⎡ ⎤⎡ ⎤
m ⎢ ẍPTX + gẋPTX ⎥ ηTX ηTXRY xPTX
⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥
⎢ m ⎥⎢ ẍPTY ⎥
+ gẋPTY ⎥ ⎢ ηTY ηTYRX ⎥⎢ xPTY ⎥
⎢ ⎥⎢ ⎢ ⎥⎢ ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥
⎢ ⎥⎢ ⎥ ⎢ ηTZ ⎥⎢ xPTZ ⎥
⎢ m ⎥⎢ ẍPTZ
⎢ + gẋPTZ ⎥
⎥ + k⎢ ⎥⎢ ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥
⎢ IXX ⎥⎢ ẍPRX + gẋPRX ⎥ ⎢ ηTYRX ηRX ⎥⎢ xPRX ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥
⎢ IYY ⎥⎢ ẍPRY + gẋPRY ⎥ ⎢η ηRY ⎥⎢ xPRY ⎥
⎣ ⎦⎢ ⎥ ⎣ TXRY ⎦⎣ ⎦
⎣ ⎦
IZZ ẍPRZ + gẋPRZ ηRZ xPRZ
, (17)
⎡ ⎤⎡ ⎤
ηTX ηTXRY xBTX
⎢ ⎥⎢ ⎥
⎢ ηTY ηTYRX ⎥⎢ xBTY ⎥
⎢ ⎥⎢ ⎥
⎢ ⎥⎢ ⎥
⎢ ηTZ ⎥⎢ xBTZ ⎥
⎢ ⎥⎢ ⎥
= k⎢ ⎥⎢ ⎥
⎢ ηTYRX ηRX ⎥⎢ xBRX ⎥
⎢ ⎥⎢ ⎥
⎢ ⎥⎢ ⎥
⎢η ηRY ⎥⎢ xBRY ⎥
⎣ TXRY ⎦⎣ ⎦
ηRZ xBRZ

where, the feedback gain g characterizes the coefficient of active damping.


Ten vibration transfer functions can be obtained from Eq. (17). Consider the RX DOF as an example. When only RX vibration is
applied to the base (xBRX ∕ = 0, xBTX = xBTY = xBTZ = xBRY = xBRZ = 0), Eq. (17) becomes
{
mẍPTY + gmẋPTY + k(ηTY xPTY + ηTYRX xPRX ) = kηTYRX xBRX
. (18)
IXX ẍPRX + gIXX ẋPRX + k(ηTYRX xPTY + ηRX xPRX ) = kηRX xBRX

The coupling between the RX input and TY ouput is clear during ηTYRX ∕ = 0. Using the Laplace transform method and solving Eq.
(18), the vibration transfer functions between “xBRX and xPRX ”, and “xBRX and xPTY ” are obtained as follows:

7
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

xPRX (s) (ms2 + gms + kηTY )kηRX − (kηTYRX )2


= , (19)
xBRX (s) (ms2 + gms + kηTY )(IXX s2 + gIXX s + kηRX ) − (kηTYRX )2

xPTY (s) (IXX s2 + gIXX s)kηTYRX


= , (20)
xBRX (s) (IXX s2 + gIXX s + kηRX )(ms2 + gms + kηTY ) − (kηTYRX )2

where, s is the Laplace variable. When the decoupling condition is satisfied, Eq. (19) can be simplified as follows and Eq. (20) equals
zero, which indicates that the RX vibration disturbance input will merely lead to the payload RX response and will not affect the
payload’s TY response.

xPRX (s) kηRX ωnRX 2


= = 2 . (21)
xBRX (s) IXX s + gIXX s + kηRX s + gs + ωnRX 2
2

Moreover, the natural frequency is given as


√̅̅̅̅̅̅̅̅̅
kηRX
ωnRX = . (22)
IXX
If the vibration disturbance input changes to the TY DOF, the same finding can be obtained for the transfer function between TY
input and TY output:

xPTY (IXX s2 + gIXX s + kηRX )kηTY − (kηTYRX )2


= . (23)
xBTY (IXX s2 + gIXX s + kηRX )(ms2 + gms + kηTY ) − (kηTYRX )2

When ηTYRX = 0, Eq. (23) becomes:

xPTY (s) kηTY ωnTY 2


= = 2 . (24)
xBTY (s) (ms + gms + kηTY ) s + gs + ωnTY 2
2

Simultaneously, the transfer function between TY input and RX output becomes zero. The same derivation and simplification are
also true for the transfer functions between TX input and RY output. Herein, the number of transfer functions was reduced to six.
To validate the effectiveness, the vibration transmissibility in the RX and TY DOFs is calculated as an example, as presented in
Fig. 6; the blue and green lines represent Eqs. (19) and (23), and the red and yellow lines represent Eqs. (21) and (24). Fig. 6 indicates
that the coupling will generate one more resonance peak in the vibration transmissibility. Taking the RX for example, derive the
frequency with the maximum magnitude via Eq. (21) and the unique solution can be found, which is the natural frequency given by
Eq. (22). Thus, the yellow line shows that only one peak occurs at the natural frequency of the decoupled RX subsystem, termed the
“natural frequency peak”. Simultaneously, there are two solutions of Eq. (19), and thereby the blue line shows two peaks: one is close
to the natural frequency peak of the decoupled RX subsystem and the other is close to the natural frequency peak of the decoupled TY
subsystem. The additional peak which is close to other subsystem can be termed the “coupling peak”. The same finding can be obtained
for TY DOF.
When the decoupling condition is met, the vibration transmissibility no longer becomes coupled, and the SUSP MIMO system can
be transformed into six SISO subsystems as shown in Fig. 3. This modeling simplification makes parameters such as resonance fre­
quencies and damping ratio manageable and the system amenable for control, which improves the system stability while guaranteeing
the performance of vibration isolation [37].

Fig. 6. Coupled and decoupled vibration transmissibility curves.

8
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

3. Controller design of the SUSP

The SUSP controller consists of force–feedback isolation and IDO steering controllers. The force–feedback isolation controller aims
to mitigate the magnitude of the vibration disturbances from the satellite at resonance frequencies. The IDO steering controller controls
the SUSP payload to steer (in the RX and RY DOFs) precisely when the reference signal is received.

3.1. Force–feedback isolation controller

The force–feedback isolation controller is similar to a sky–hook damper [38,39], which generates a control force proportional to the
absolute acceleration feedback of the payload. [40] stated that force feedback brings more stability to the control system than ac­
celeration feedback when the payload is flexible. In real situation, the payload structure is complicated and cannot be regarded as a
rigid body. Consequently, the strut axial force Fl was selected as the feedback to generate the SUSP control force Faiso according to Eq.
(6). The isolation controller diagram is shown in Fig. 7. Becauce the isolation actuation force faisoi generated by the ith strut (i = 1,2,⋯,
6) is only concerned with the axial force fli measured by the force sensor mounted on the same strut, the isolation controller can be
regarded as a part of the strut system.
The force–feedback gain is given by g = 2ξωc , where ξ is the damping ratio and ωc is the corner frequency of the transfer functions.
After the decoupling design, the 6–DOF transfer functions of the SUSP have the same form as in Eqs. (21) and (24). Meanwhile, the
natural frequency matrix Ωn is easy to calculate using the parameters in Table 1, and the same equation forms as Eq. (22):
Ωn = diag{ωnTX , ⋯, ωnRZ } = 2π ⋅ diag{1.01, 1.01, 1.43, 1.47, 1.47, 2.02} (25)

It can be observed that the 6–DOF natural frequencies are approximately equal. To achieve an approximate critical damping in 6–DOF
√̅̅̅
vibration transmissibility, let ξ = 1 / 2 and ωc = ωnTZ ; ωnTZ is close to the average of the 6–DOF natural frequencies according to Eq.
(25).

3.2. IDO steering controller

The IDO steering controller is designed to compensate for the jitter caused by a large payload mass ratio and steer the payload
precisely and stably. Owing to the decoupling design, the SUSP is transformed into six single–DOF SISO subsystems, which demon­
strates that introducing the control force in one DOF can separately drive the payload to steer in the same DOF. Accordingly, the
distributed IDO steering controllers for the RX and RY subsystems can be designed separately. The IDO steering controller of the SUSP
is presented in Fig. 8, which includes a PID controller (the black box) for tracking the reference displacement xRi (i = RX,RY), and an
improved disturbance observer (the blue and red boxes) for eliminating the jitter caused by the large payload mass ratio. As well,
parameter tuning can be implemented separately for RX and RY subsystems to obtain a better steering performance.
The PID controller is typically given by
ui (s) kIi
GCi (s) = = kPi + + kDi s, i = RX, RY, (26)
ei (s) s

where kPi , kIi and kDi are the feedback gains acting on the proportional, integral and differential tracking errors, respectively, ei is the
tracking error, and ui is the PID output.
To eliminate the jitter, the disturbance observer is introduced and the jitter is recognized as a disturbance state variable di of the RX
and RY subsystems. The blue box includes a disturbance observer based on the nominal transfer functions Gni (s) for the actual RX and
RY subsystems GPi (s). To derive Gni (s), assume that FPstr = (0, 0, 0, fPstrRX , fPstrRY , 0)T and XB = 0, and Eq.(4) becomes

Fig. 7. Strut system diagram.

9
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

Fig. 8. Block diagram of the IDO steering controller.

MP ẌP + gMP ẊP + kJTP JP XP = FPstr . (27)

Correspondingly, Gni (s) is derived as follows:


xPi (s) 1 1
Gni (s) = = ⋅ , i = RX, RY. (28)
fPstri (s) Ii s2 + gs + ωni

d i = (εi +di )GPi (s)G−ni1 (s) − εi ≈ di is the estimated value of di . To reduce the effects of the measurement noise of xPi , a
Therefore, ̂
second–order low–pass filter Q(s) (σ1 = 1) is introduced into the disturbance observer.
The transmissibility of ui and xPi is given as
GPi (s)Gni (s)
Guxi (s) = , (29)
Q(s)GPi (s) + [1 − Q(s)]Gni (s)

and that of di and xPi is given as


GPi (s)Gni (s)[1 − Q(s)]
Gdxi (s) = . (30)
Q(s)GPi (s) + [1 − Q(s)]Gni (s)

Note that s → 0, Q(s) = 1, Guxi (s) = Gni (s), and Gdxi (s) = 0; when s → ∞, Q(s) → 0, Guxi (s) = GPi (s), and Gdxi (s) = GPi (s). This implies
that the cut–off frequency of Q(s) determines the effective bandwidth of the disturbance observer.
Assuming that GPi (s) = Gni (s)[1 + Δ(s)], where Δ(s) is the error term, the system is robust on the condition [41]
‖ Δ(s)T(s) ‖∞ = ‖ Δ(s)Q(s) ‖∞ ≤ 1. (31)

where T(s) = 1 − S(s) is the complementary sensitivity function and the sensitivity function is S(s) = 1 − Q(s). Owing to this condition,
as shown in Eq.(31), the effective bandwidth of the disturbance observer is limited.
To obtain a higher gain in the effective bandwidth, an additional loop was designed to improve the disturbance observer, as shown
in the red box in Fig. 8. Eqs. (29) and (30) become

Fig. 9. Configuration of SUSP struts.

10
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

GPi (s)Gni (s)


Guxi (s) = , (32)
σ1 Q(s)GPi (s) + [1 − (σ 1 + σ2 )Q(s)]Gni (s)

and
GPi (s)Gni (s)[1 − (σ1 + σ 2 )Q(s)]
Gdxi (s) = , respectively, (33)
σ1 Q(s)GPi (s) + [1 − (σ 1 + σ2 )Q(s)]Gni (s)

where σ1 + σ2 = 1 maintains the attenuation of di . Condition 0 ≤ σ1 , σ2 ≤ 1 guarantees that the low–frequency gain in Eq. (32) is
higher than that in Eq.(29), which implies a wider low–frequency band of vibration suppression and higher accuracy of payload
steering. The advantage of the decoupling design is that a distributed controller is possible for each subsystem. The performance of the
SUSP in “working DOFs” (RX and RY DOFs in this study) can be improved by merely applying the distributed controller to the con­
cerned subsystems, which reduces the measurement resources and control difficulty, and the remaining DOFs are not affected.

4. Dynamic model of the SUSP

The rigid–flexible coupling dynamic model of the SUSP is developed and analyzed in LMS Virtual.Lab software. The model includes
flexible and rigid parts: the membrane and flexible joint in each strut are flexible, and the remaining parts such as the payload and the
satellite are rigid.
To develop the strut model according to the configuration shown in Fig. 9, the rigid parts mainly include a voice–coil motor (VCM)
[42] with a magnet fixed to the BP to reduce the impact of the strut local modes [43] and a voice coil, and a rod. And the flexible parts
are the membrane and the flexible joint. Using the multipoint constraint (MPC), the outer circle of the membrane was connected to the
BP, the inner circle of the membrane and the lower end of the flexible joint were connected to the rod’s two ends, and the flexible
joint’s upper end was connected to the MP.
The membrane provides k for the strut and acts as the lower joint by cutting out five three–symmetric gaps on the sheet Beryllium
bronze alloy (QBe2) material. And the flexible joint acts as the upper joint by cutting out a pair of elliptical gaps on the cylindrical
titanium alloy (TC4) material. The material structural properties are listed in Table 2. In this study, the gaps of the membrane and
flexible joint were carefully designed to make the rotational stiffness of the two joints to be much smaller than the axial stiffness to
avoid additional impacts. And then the stiffness is checked through the FEA and k is consistent with Table 1, as shown in Fig. 10. Next,
the Craig–Bampton modes of the membrane and flexible joint were calculated and attached to the strut model, which is a common
method to build a rigid–flexible coupling model.
To build the model of whole system, the struts were placed according to the architecture parameters listed in Table 1 with the
centers of the membranes at the location of b1 ∼ b6 and flexible joints at p1 ∼ p6 , and the bottoms of the struts are fixed to the BP. Then,
the BP is fixed to the satellite (mB = 1556 kg, IXXB = 1307 kg × m2, IYYB = 1301 kg × m2, IZZB = 1768 kg × m2). And the MP was fixed
to the payload.
At present, the system model is developed. The structural modes of the SUSP in 6 DOFs can be analyzed and displayed in Fig. 11 (the
satellite and other rigid parts are hidden), and the modal frequencies are identical to the natural frequency matrix Ωn . The vibration
shape diagram shows that the RX and RY rotation axes of the SUSP coincide with the optimal CP designed in Section 2.2, as shown in
Fig. 11(d) and (e), respectively, which demonstrates the effectiveness of the decoupling design. In Section 5, this model is used to
implement the simulations to verify the effectiveness of decoupling and SUSP functions.

5. SUSP simulations

This section introduces the simulations of SUSP using the controller designed in Section 3 and the rigid–flexible coupling dynamic
model established in Section 4. The simulations are divided into two stages: 1) the first stage is vibration isolation with random
micro–vibration excitation attached to the satellite, demonstrating the effectiveness of the decoupling design and vibration isolation;
2) the second stage is the isolation–steering integrated control, which implements the 0.1◦ circular tracing under the same vibration
condition to verify the vibration isolation and precise steering functions of the SUSP. In this study, the control performance is validated
by the 6–DOF vibration transmissibility between XB and XP . Since the influence of satellite attitude control on the vibration control is
limited, the satellite attitude control is neglected in simulations. The simulations are implemented with LMS Virtual.Lab Motion
software for running the rigid–flexible coupling dynamic model and Matlab Simulink software for running the control algorithm.

Table 2
Material structural properties.
Items QBe2 TC4
11 − 2 11 − 2
Young modulus 1.3 × 10 N × m 1.1 × 10 N × m
Poisson ratio 0.30 0.33
Density 8250 kg × m − 3 4400 kg × m − 3

11
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

Fig. 10. FEA of (a) the membrane and (b) flexible joint.

Fig. 11. Structural modes of the SUSP in 6 DOFs, (a) TX (1.05 Hz), (b) TY (1.04 Hz), (c) TZ (1.45 Hz), (d) RX (1.50 Hz), (e) RY (1.49 Hz), and (f) RZ
(2.04 Hz).

5.1. Simulation of vibration isolation and 6–DOF decoupling verification

The simulation in this stage was performed using only a vibration isolation controller. First, the isolation controller is an open–loop
(g = 0), and the acceleration sensors are attached to the origins of {P} and {B} to measure the accelerations of the payload and
satellite. The sensor data are processed, and the 6–DOF open–loop vibration transmissibility between XB and XP is shown in Fig. 12.
Two versions of the SUSP were examined, i.e., coupled and decoupled SUSP. It can be observed that the coupling causes coupling
peaks in TX, TY, RX, and RY, as shown in Fig. 12(a), whereas Fig. 12(b) indicates that the decoupling design can significantly attenuate
the coupling peaks. Furthermore, the coupled transmissibility between the TY input and RX output of the two versions is presented in
Fig. 13. Although there is residual coupled vibration response of the decoupled version (red line) cause by the small movement of SUSP
struts, it is approximately − 20 dB lower than that of the coupled version (blue line). The results indicate the effectiveness of the

12
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

Fig. 12. 6–DOF open–loop vibration transmissibility of two SUSP versions. (a) Coupled and (b) decoupled.

Fig. 13. Open–loop coupled transmissibility between TY and RX of two versions.

13
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

decoupling design in limiting the coupling influence. Therefore, the SUSP can be regarded as approximately six SISO subsystems after
the decoupling design.
The SUSP was decoupled into six subsystems with a single DOF. To verify the vibration control performance, a closed–loop
simulation (g = 2ξωnTZ ) was performed and the 6–DOF closed–loop vibration transmissibility is shown in Fig. 14. The result exhibits a
roll–off rate of − 40 dB/decade after the corner frequencies and no amplification of magnitude at full band, which illustrates that the
force feedback gain is suitable, and the SUSP can isolate 6–DOF vibration disturbances after the natural frequencies of each subsystem.

5.2. Simulation of isolation–steering integrated control

In this stage, the isolation and steering controllers were included, where the IDO steering controller was applied in three steps to
achieve a 0.5 Hz, 0.1◦ circular–trace steering for comparison. The first step involved steering using only a PID controller. The
simulation results are shown in Fig. 15, where the blue line is the trace, and the red line is the tracking error within 2′′ . The PID
controller cannot eliminate the low–frequency jitter caused by the large payload mass ratio, as shown in the “tracking error” diagram
in Fig. 15.
In the second step, a disturbance observer was introduced to reduce the jitter. Considering RX as an example, the transmissibility of
the vibration response GPRX (s) was obtained and plotted and the nominal transfer function GnRX (s) was shown in Fig. 16(a). Subse­
quently, 1/Δ(s) was calculated, and it is displayed in Fig. 16(b). According to Eq. (31), the second–order low–pass filter Q(s) with a
cut–off frequency of 20 Hz is feasible. The same process of designing the disturbance observer was applied for RY, and the results
demonstrate the effectiveness of the disturbance observer; the jitter is compensated and the error of 1.22′′ remains (Fig. 17).
In the next step, the IDO was introduced and parameters σ1 and σ 2 were tuned to achieve a higher tracking accuracy. The results in
Fig. 18 show that the error was further reduced to 0.13′′ , compared to Fig. 17 in the second step, verifying the capacity to eliminate the
jitter caused by the large payload mass ratio and improve the steering accuracy affected by the vibration. In addition, the 6–DOF
vibration transmissibility was calculated, as shown in Fig. 19. Compared with Fig. 14, the magnitudes of RX and RY in the low­
–frequency bandwidth are reduced to below 0 dB, except for the steering frequency of 0.5 Hz. Meanwhile, diagrams for other DOFs are
close to Fig. 14, which indicates that after the decoupling design, the distributed control can separately improve the vibration control
performance (such control gain and bandwidth) for the concerned DOF. In addition, the maximum values of the stroke and actuation
force of the SUSP struts are 1.5 mm and 40.56 N, respectively, which are all within the allowable ranges of the chosen VCM.

6. Conclusion

This study is concerned with the Stewart platform for in–orbit vibration isolation and precise steering, and designs a 6–DOF
decoupled SUSP with a large payload mass ratio. First, the system numerical model was investigated, and the decoupling condition was
derived. On this basis, the vibration transfer functions can be simplified, and the SUSP is designed and decoupled into six single–DOF
SISO subsystems with the same form of the transfer function. This modeling simplification facilitates controller design and improves
control performance. The natural frequency matrix can be easily designed, and the feedback gain of the vibration isolation controller
for each subsystem can be identical. In addition, a distributed IDO steering controller based on the nominal system simplified by the
decoupling design for the RX and RY subsystems was developed. Considering the impact of the large payload mass ratio as disturbances
attached to the two subsystems, a disturbance observer was introduced, and its low–frequency performance was subsequently
improved. After rigid–flexible dynamic modeling, two stages of simulation were performed, i.e., the vibration isolation and

Fig. 14. 6–DOF close–loop vibration transmissibility of isolation control.

14
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

Fig. 15. 0.1◦ , 0.5 Hz, 60 s, circular trace steering (PID).

Fig. 16. Design of the disturbance observer. (a) Nominal transfer function GnRX (s) and (b) low–pass filterQ(s).

Fig. 17. 0.1◦ , 0.5 Hz, 60 s, circular trace steering (PID and disturbance observer).

isolation–steering integrated stages. The vibration isolation stage results indicate a reduction of approximately − 20 dB in the coupled
responses and a − 40 dB/decade roll–off rate of the 6–DOF vibration transmissibility; the isolation–steering integrated stage results
demonstrate full–band vibration isolation and high steering precision in the concerned RX and RY DOFs, which illustrates the
effectiveness of the decoupling design and isolation–steering integrated controller. In addition, the decoupled system allows different
control algorithms to be applied to different subsystems to achieve a higher control performance. In the future when conditions are
available, the experiment can be implemented of further verify the effectiveness of the decoupling method and the vibration isolation

15
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

Fig. 18. 0.1◦ , 0.5 Hz, 60 s, circular trace steering (IDO steering controller).

Fig. 19. 6–DOF close–loop vibration transmissibility of isolation–steering integrated control.

and steering functions of the SUSP.

CRediT authorship contribution statement

Chao Liang: Conceptualization, Investigation, Methodology, Formal analysis, Writing – original draft, Writing – review & editing.
Weipeng Li: Resources, Funding acquisition, Supervision, Writing – review & editing, Project administration. Hai Huang: Project
administration, Supervision, Methodology. Zhuokun Wu: Formal analysis, Writing – review & editing.

Declaration of Competing Interest

The authors declare the following financial interests/personal relationships which may be consideredas potential competing
interests.

Data availability

All data generated or analysed during this study are included in this published article.

16
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

Acknowledgement

This research was supported by the National Key R&D Program of China (grant no. 2021YFA1003503).

Appendix A

Let the deviation from the Z–axis be (δx , δy , 0)T ; the first and second strut components in JP are modified as follows:
( )T
1 1
p1 = rp cos αp − δx , rp sin αp − δy , pz1
2 2
( )T , (A.1)
1 1
p2 = rp cos αp − δx , − rp sin αp − δy , pz1
2 2

s1 and s2 remain unchanged. Then, the modification of GJ can be expressed as


⎡ ⎤
ηTX ηTXRY ηTXRZ
⎢ ηTY ηTYRX ηTYRZ ⎥
⎢ ⎥
⎢ ηTZ ηTZRX ηTZRY ⎥
GJ = ⎢

⎥, (A.2)
⎢ ηTYRX ηTZRX ηRX ηRXRY ηRXRZ ⎥

⎣η ηTZRY ηRXRY ηRY ηRYRZ ⎦
TXRY
ηTXRZ ηTYRZ ηRXRZ ηRYRZ ηRZ

where, ηTX , ηTY , ηTZ , ηTYRX , and ηTXRY are the same as those in Eq. (13), and the remaining transfer coefficients are
[ ( )]
1 ( )
ηRX = 3pz1 sinθ pz1 sinθ − 2rp cosθcos αp − φ + 3cos2 θ rp 2 + 2δy 2 (A.3)
2
[ ( )]
1 ( )
ηRY = 3pz1 sinθ pz1 sinθ − 2rp cosθcos αp − φ + 3cos2 θ rp 2 + 2δx 2 (A.4)
2
[ ( ) ]
1 ( )
ηRZ = 3sin2 θ 2rp 2 sin2 αp − φ + δx 2 + δy 2 (A.5)
2

ηRXRY = − 6δx δy cos2 θ (A.6)


[ ( )]
1
ηRXRZ = 3δx sinθ pz1 sinθ − rp cosθcos αp − φ (A.7)
2
[ ( )]
1
ηRYRZ = 3δy sinθ pz1 sinθ − rp cosθcos αp − φ (A.8)
2

ηTXRZ = 3δy sin2 θ (A.9)

ηTYRZ = − 3δx sin2 θ (A.10)

ηTZRX = − 6δy cos2 θ (A.11)

ηTZRY = 6δx cos2 θ (A.12)


Compared with Eq. (13), the deviation will cause more coupled transfer coefficients in GJ , implying a more complicated system
coupling. When the decoupling condition given by Eqs. (15) and (16) is satisfied, ηTYRX , ηTXRY , ηRXRZ , and ηRYRZ equal 0, while the
remaining coupled transfer coefficients, which are linear with the deviation, remain unchanged; this indicates the system is still
coupled. To evaluate the degree of coupling, the coupling ratio index (CRI) is defined by:
ηij
CRIij = √̅̅̅̅̅̅̅, (A.13)
ηi ηj

where, i, j = TX, TY, TZ, RX, RY, RZ and i ∕


= j. GJ is diagonal if all the CRIs are equal to zero. When CP is not on the Z–axis in real
situations, the system can still be regarded as approximately decoupled if the CRI is small.

17
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

References

[1] Y. Xu, H. Liao, L. Liu, Y. Wang, Modeling and robust H–infinite control of a novel non–contact ultra–quiet Stewart spacecraft, Acta Astronaut 107 (2015)
274–289, https://doi.org/10.1016/j.actaastro.2014.11.033.
[2] E.H. Anderson, J.P. Fumo, R.S. Erwin, Satellite ultraquiet isolation technology experiment (SUITE), in: Aerosp. Conf. Proceedings, 2000 IEEE, 2000,
pp. 299–313, https://doi.org/10.1109/AERO.2000.878441.
[3] X. Sun, B. Yang, L. Zhao, X. Sun, Optimal design and experimental analyses of a new micro–vibration control payload–platform, J. Sound Vib. 374 (2016) 43–60,
https://doi.org/10.1016/j.jsv.2016.04.007.
[4] J. Spanos, Z. Rahman, G. Blackwood, Soft 6–axis active vibration isolator, in: Proc. Am. Control Conf, 1995, https://doi.org/10.1109/acc.1995.529280.
[5] A.J. Bronowicki, Vibration isolator for large space telescopes, J. Spacecr. Rockets. 43 (2006) 45–53, https://doi.org/10.2514/1.12036.
[6] E.H. Anderson, D.J. Leo, M.D. Holcomb, Ultraquiet platform for active vibration isolation, in: I. Chopra (Ed.), Proc. SPIE – Int. Soc. Opt. Eng., Ed., 1996,
pp. 436–451, https://doi.org/10.1117/12.239046.
[7] D. Stewart, A platform with six degrees of freedom, Proc. Inst. Mech. Eng. 180 (1965) 371–386, https://doi.org/10.1243/PIME_PROC_1965_180_029_02.
[8] M. Furqan, M. Suhaib, N. Ahmad, Studies on stewart platform manipulator: a review, J. Mech. Sci. Technol. 31 (2017) 4459–4470, https://doi.org/10.1007/
s12206-017-0846-1.
[9] A. Preumont, M. Horodinca, I. Romanescu, B. de Marneffe, M. Avraam, A. Deraemaeker, F. Bossens, A. Abu Hanieh, A six–axis single–stage active vibration
isolator based on Stewart platform, J. Sound Vib. 300 (2007) 644–661, https://doi.org/10.1016/j.jsv.2006.07.050.
[10] Dae–Oen Lee, G. Park, J.–.H. Han, Hybrid isolation of micro vibrations induced by reaction wheels, J. Sound Vib. 363 (2016) 1–17, https://doi.org/10.1016/j.
jsv.2015.10.023.
[11] W. Chi, D. Cao, D. Wang, J. Tang, Y. Nie, W. Huang, Design and experimental study of a VCM–based stewart parallel mechanism used for active vibration
isolation, Energies 8 (2015) 8001–8019, https://doi.org/10.3390/en8088001.
[12] V. Babuška, R.S. Erwin, L.A. Sullivan, System identification of the suite isolation platform: comparison of ground and flight experiments, in: 44th AIAA/ASME/
ASCE/AHS/ASC Struct. Struct. Dyn. Mater. Conf., 2003, pp. 1–11.
[13] A. Joshi, W. Kim, System Identification and Multivariable Control Design for a Satellite UltraQuiet Isolation Technology Experiment (SUITE), Eur. J. Control. 10
(2004) 174–186.
[14] J.H. Jacobs, J.A. Ross, S. Hadden, M. Gonzalez, Z. Rogers, B.K. Henderson, Miniature vibration isolation system for space applications: phase II, in: E.
H. Anderson (Ed.), Ind. Commer. Appl. Smart Struct. Technol, Ed., 2004, p. 32, https://doi.org/10.1117/12.540007.
[15] M.B. McMickell, T. Kreider, E. Hansen, T. Davis, M. Gonzalez, Optical payload isolation using the miniature vibration isolation system (MVIS–II, in: L.P. Davis,
B.K. Henderson, M.B. McMickell (Eds.), Ind. Commer. Appl. Smart Struct. Technol., Eds., 2007, 652703, https://doi.org/10.1117/12.715446.
[16] V. Camelo, A. Bronowicki, R. Hejal, S. Simonian, S. Brennan, Damping and isolation concepts for vibration suppression and pointing performance, in: 50th
AIAA/ASME/ASCE/AHS/ASC Struct. Struct. Dyn. Mater. Conf., American Institute of Aeronautics and Astronautics, Reston, Virigina, 2009, pp. 1–16, https://
doi.org/10.2514/6.2009-2637.
[17] R.G. Cobb, J.M. Sullivan, A. Das, L.P. Davis, T.T. Hyde, T. Davis, Z.H. Rahman, J.T. Spanos, Vibration isolation and suppression system for precision payloads in
space, Smart Mater. Struct. 8 (1999) 798–812, https://doi.org/10.1088/0964-1726/8/6/309.
[18] L. Tang, X. Guan, D. Zhang, Y. He, Control Design of Ultra–Quiet Spacecraft Platform, IFAC, 2013, https://doi.org/10.3182/20130902-5-de-2040.00062.
[19] L. Tang, Z. Guo, X. Guan, Y. Wang, K. Zhang, Integrated control method for spacecraft considering the flexibility of the spacecraft bus, Acta Astronaut 167
(2020) 73–84, https://doi.org/10.1016/j.actaastro.2019.08.030.
[20] L. Tang, K. Zhang, X. Guan, R. Hao, Y. Wang, Dynamic modeling and multi–stage integrated control method of ultra–quiet spacecraft, Adv. Sp. Res. 65 (2020)
271–284, https://doi.org/10.1016/j.asr.2019.09.001.
[21] J.E. McInroy, J.F. Obrien, A.A. Allais, Designing micromanipulation systems for decoupled dynamics and control, IEEE/ASME Trans. Mechatronics. 20 (2015)
553–563, https://doi.org/10.1109/TMECH.2013.2296154.
[22] Z. Tong, C. Gosselin, H. Jiang, Dynamic decoupling analysis and experiment based on a class of modified Gough–Stewart parallel manipulators with line
orthogonality, Mech. Mach. Theory. 143 (2020), 103636, https://doi.org/10.1016/j.mechmachtheory.2019.103636.
[23] X.L. Yang, H.T. Wu, B. Chen, S.Z. Kang, S.L. Cheng, Dynamic modeling and decoupled control of a flexible Stewart platform for vibration isolation, J. Sound Vib.
439 (2019) 398–412, https://doi.org/10.1016/j.jsv.2018.10.007.
[24] C. Yang, Z. Qu, J. Han, Decoupled–space control and experimental evaluation of spatial electrohydraulic robotic manipulators using singular value
decomposition algorithms, IEEE Trans. Ind. Electron. 61 (2014) 3427–3438, https://doi.org/10.1109/TIE.2013.2278958.
[25] Y. Yi, J.E. McInroy, F. Jafari, Generating classes of Orthogonal Gough–Stewart Platforms, in: Proc. – IEEE Int. Conf. Robot. Autom, 2004, pp. 4969–4974,
https://doi.org/10.1109/robot.2004.1302505, 2004.
[26] H.Z. Jiang, J.F. He, Z.Z. Tong, W. Wang, Dynamic isotropic design for modified Gough–Stewart platforms lying on a pair of circular hyperboloids, Mech. Mach.
Theory. 46 (2011) 1301–1315, https://doi.org/10.1016/j.mechmachtheory.2011.04.003.
[27] J.E. McInroy, J.C. Hamann, Design and control of flexure jointed hexapods, IEEE Trans. Robot. Autom. 16 (2000) 372–381, https://doi.org/10.1109/
70.864229.
[28] J.E. McInroy, F. Jafari, Finding symmetric orthogonal Gough–Stewart platforms, IEEE Trans. Robot. 22 (2006) 880–889, https://doi.org/10.1109/
TRO.2006.878975.
[29] H.Z. Jiang, Z.Z. Tong, J.F. He, Dynamic isotropic design of a class of Gough–Stewart parallel manipulators lying on a circular hyperboloid of one sheet, Mech.
Mach. Theory. 46 (2011) 358–374, https://doi.org/10.1016/j.mechmachtheory.2010.10.008.
[30] M. Viso, et al., Spitzer space telescope, in: M. Gargaud, W.M. Irvine, R. Amils, H.J. (Jim) Cleaves, D.L. Pinti, J.C. Quintanilla, et al. (Eds.), Encycl. Astrobiol.,
Eds., Springer Berlin Heidelberg, Berlin, Heidelberg, 2015, pp. 2325–2329, https://doi.org/10.1007/978-3-662-44185-5_1492.
[31] M.A. Greenhouse, The james webb space telescope: mission overview and status, AIAA Sp. Conf. Expo. (2012) 1–10, https://doi.org/10.2514/6.2012-5100,
2012.
[32] W.P. Li, B. Luo, H. Huang, Active vibration control of flexible joint manipulator using input shaping and adaptive parameter auto disturbance rejection
controller, J. Sound Vib. 363 (2016) 97–125, https://doi.org/10.1016/j.jsv.2015.11.002.
[33] M. Li, Y. Zhang, Y. Wang, Q. Hu, R. Qi, The pointing and vibration isolation integrated control method for optical payload, J. Sound Vib. 438 (2019) 441–456,
https://doi.org/10.1016/j.jsv.2018.09.038.
[34] Y. Kong, H. Huang, Vibration isolation and dual–stage actuation pointing system for space precision payloads, Acta Astronaut 143 (2018) 183–192, https://doi.
org/10.1016/j.actaastro.2017.11.038.
[35] Y. Kong, H. Huang, Performance enhancement of disturbance–free payload with a novel design of architecture and control, Acta Astronaut 159 (2019) 238–249,
https://doi.org/10.1016/j.actaastro.2019.03.061.
[36] A.A Hanieh, Active structure Laboratory, ULB, PH. D. Thesis, 2003.
[37] P. Mukherjee, B. Dasgupta, A.K. Mallik, Dynamic stability index and vibration analysis of a flexible Stewart platform, J. Sound Vib. 307 (2007) 495–512,
https://doi.org/10.1016/j.jsv.2007.05.036.
[38] D.C. Karnopp, A.K. Trikha, Comparative study of optimization techniques for shock and vibration isolation, Transactions of the ASME, J. Eng. Industry, Series B
91 (1969) 1128–1132.
[39] C.E. Kaplow, J.R. Velman, Active local vibration isolation applied to a flexible space telescope, AIAA J. Guidance and Control 3 (1980) 227–233.
[40] A. PREUMONT, A. FRANÇOIS, F. BOSSENS, A. ABU–HANIEH, Force feedback versus acceleration feedback in active vibration isolation, J. Sound Vib. 257
(2002) 605–613, https://doi.org/10.1006/jsvi.2002.5047.

18
C. Liang et al. Journal of Sound and Vibration 553 (2023) 117668

[41] D. John, F. Bruce, T. Allen, Feedback control Theory, (1990) 50–52.


[42] H.J. Chen, R.M. Bishop, B.N. Agrawal, Payload pointing and active vibration isolation using hexapod platforms, in: 44th AIAA/ASME/ASCE/AHS/ASC Struct.
Struct. Dyn. Mater. Conf., Reston, Virigina, American Institute of Aeronautics and Astronautics, 2003, pp. 2196–2214, https://doi.org/10.2514/6.2003-1643.
[43] W. Li, H. Huang, X. Zhou, X. Zheng, Y. Bai, Design and experiments of an active isolator for satellite micro–vibration, Chin. J. Aeronaut. 27 (2014) 1461–1468,
https://doi.org/10.1016/j.cja.2014.10.012.

19

You might also like