You are on page 1of 30

Mechanism and Machine Theory 172 (2022) 104770

Contents lists available at ScienceDirect

Mechanism and Machine Theory


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

A design approach for gravity compensators using planar four-bar


mechanisms and a linear spring
Vu Linh Nguyen
Department of Mechanical Engineering, National Chin-Yi University of Technology, Taichung 411030, Taiwan

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

Keywords: This paper presents a design approach for gravity compensators using planar four-bar mecha­
Gravity compensation nisms and a linear spring. This work enables the development of a class of gravity compensators
Static balancing encompassing 42 different types, which are characterized by high performance and kinematic
Planar four-bar mechanism
simplicity. The gravity compensators are constructed by combining planar four-bar mechanisms
Spring mechanism
Optimal design
with a rotating mass and attaching a linear spring to each mechanism, then permuting the springs.
The parameters of the gravity compensators are derived from an optimization procedure that
minimizes the actuator torque within a specified balancing zone. The performances of the gravity
compensators were demonstrated via both numerical examples and experiments. The results
showed their design feasibilities and high performances in which a torque reduction rate of 87.8%
was practically achieved. Lastly, an application of the proposed gravity compensators to serial
robots is described. It was found that the actuator torque of a serial robot over a prescribed
workspace could theoretically be reduced by 98.2%. A prototype of a serial robot was also built to
validate the applicability of the gravity compensators.

1. Introduction

Gravity compensation of mechanisms has a substantial technological significance in mechanism and robot design [1–4]. Complete
gravity compensation for a mechanism enables it to be stabilized at any position in its workspace with zero input forces and torques.
The input forces and torques required to maintain a mechanism are partially eliminated with incomplete gravity compensation. With
reductions in the input forces and torques, gravity compensation for a mechanism permits it to operate with lesser energy con­
sumption, smaller actuator sizes, increased safety, and improved dynamic responses [3–5].
Counterweight and spring methods are the most popular ways to implement gravity compensation for mechanisms [2,6]. In
counterweight methods, additional weights are attached to a mechanism so that the total mass center of the entire system can be
maintained during operations [7–11]. Spring methods exploit the elastic energy stored by springs to eliminate the variation of the
potential energy of a mechanism [12–16]. Compared to counterweight methods, their spring counterparts usually offer a lighter
mechanical structure and lesser inertial forces [3,4]. With the above-described advantages, spring methods are preferred in many
situations, especially where a lightweight mechanism is prioritized [17–19].
The gravity compensation design of a one-degree-of-freedom (1-DoF) rotating link with a mass is the simplest design in this field
[1–4]. When a spring method is adopted for gravity compensation, the springs are articulated to a rotating link via an auxiliary
mechanism to generate nonlinear elastic forces to oppose the gravitational force of the rotating link [12–16]. Such a spring mechanism

E-mail address: vlnguyen@ncut.edu.tw.

https://doi.org/10.1016/j.mechmachtheory.2022.104770
Received 17 November 2021; Received in revised form 10 January 2022; Accepted 1 February 2022
Available online 2 March 2022
0094-114X/© 2022 Elsevier Ltd. All rights reserved.
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Nomenclature

d0 Spring length at the initial position


d1 Spring length at an instantaneous position
Fai Actuator force vector on link i
Fij Reaction force vector on link i
Fik External force vector on link i
Fs Elastic force vector
g Gravitational acceleration vector
k Spring stiffness coefficient
li Length of link i
lci Position of the mass center of robot link i + 1
L0 Free length of the spring
LCi Distance from the accumulated mass center to joint Oi
Li Length of robot link i + 1
mi Mass of link i
mc Mass of the rotating link
mci Mass of robot link i + 1
MCi Accumulated rotating mass applying to joint Oi
rci Position vector associated with the mass center of link i
rij Position vector associated with the reaction force on link i
rik Position vector associated with the external force on link i
s0 Initial spring extension
Ta Actuator torque vector exerted on link i
Tri Reaction torque vector exerted on link i
Ta4 Compensated torque
Tun Uncompensated torque
X Design variable vector of type E
Y Design variable vector of type F
α Joint angle of link 2
θ Joint angle of the rotating link
θin Initial position of the rotating link
θen End position of the rotating link
δt Torque reduction rate (TRR)
ΩX Wrench vector of type E
ΨX Wrench vector of type F
ρs Extension-to-length ratio (ETLR)
ζlj Lower bound of variable Xj
ζuj Upper bound of variable Xj
μs Unit vector along the spring’s longitudinal axis

for gravity compensation can be called a gravity compensator. Numerous researchers have endeavored to render well-designed gravity
compensators and then exploit them for the gravity compensation of robotic manipulators [20–23], assistive devices [24–27], and
surgical devices [28,29]. For instance, Rahman et al. [30] used a pulley-cable mechanism with a linear spring to make a gravity
compensator and then extended it to the gravity compensation of articulated mechanisms. Herder [6] conceptualized a
zero-free-length (ZFL) spring and proposed several practical embodiments for it in building gravity compensators. Ulrich and Kumar
[31] presented a gravity compensator based on a noncircular pulley mechanism with linear springs and applied it to articulated
manipulators. Fedorov and Birglen [32] designed a gravity compensator using a pair of differential noncircular pulleys with a
constant-length cable and linear springs. Arakelian and Zhang [33] presented a gravity compensator based on an inverted slider-crank
mechanism with a linear spring. Nguyen et al. [34] used a geared five-bar mechanism and a linear spring to build a gravity
compensator and then applied it to planar articulated robotic arms. Koser [35] exploited a cam-follower mechanism with a linear
spring to develop a gravity compensator and showed its application to serial robotic arms. Tschiersky et al. [24] proposed a gravity
compensator based on a flexure spring to design an assistive elbow orthosis. Gravity compensators were also built by adopting other
spring mechanisms, such as a coaxial gear train with noncircular gears and torsion bars [36], cam-follower mechanisms with linear
springs [37,38], a Cardan gear mechanism with a linear spring [39], slider-crank mechanisms with linear springs [23,40], a Scotch
yoke mechanism with a linear spring [41], a bidirectional spiral pulley negative stiffness mechanism with linear springs [42], a
multi-winding wire-driven mechanism with a linear spring [18], and a four-bar compliant mechanism with torsion springs [43].
Furthermore, passive constant-force mechanisms (CFMs), whose output forces are nearly constant over a prescribed deflection range,
can also be used as gravity compensators [44–50]. Rigid-link CFMs are usually constructed with a rigid-link mechanism (e.g., a five-bar

2
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 1. Construction process of the gravity compensators using planar four-bar mechanisms and a linear spring.

mechanism [44], a rotational cam-slider mechanism [45], and a double cam-follower mechanism [46]) and mechanical springs.
Compliant CFMs are realized from the deformation of their flexible members, such as the utilization of the second buckling mode of
flexible beams [47], the combination of a multi-layer flexure and bistable beam [48], and the use of folding beam and bistable beam
mechanisms [49]. A survey of the design and modeling of CFMs has been presented by Wang and Xu [50].
The literature, as reviewed above, has so far exhibited no convergence on spring mechanisms in the construction of a gravity
compensator. Although current gravity compensators are the result of progressive developments, limitations in each design still
inevitably exist. For example, the use of ZFL springs in gravity compensators [6,30] may lead to practical issues related to friction and

3
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 2. Six different types of gravity compensators in Group I.

dimensional errors. Noncircular pulley mechanisms [31,32,42] may require a large volume to embed the pulleys and springs and suffer
from friction on the contact between the pulleys and cables. An inverted slider-crank mechanism [33] requires a large volume for the
spring implementation and may be affected by friction on the sliding motion. The adoption of gear trains for motion/force trans­
missions in gravity compensators [34,36,39,41] may cause structural complexity and cumbersomeness, transmission errors because of
the gear backlash, and power losses due to the gear friction. Cam-follower mechanisms [35,37,38] may introduce an inadequate force
transmission when even minor dimensional errors on the parts occur or when the contact force between the cam and follower is too
large. Slider-crank mechanisms [23,40] require a large spring preload that causes inconvenience for the spring installation. Compliant
mechanisms [24,43] are limited in bearing large loads and range of motion. The above-specified limitations may challenge the existing
gravity compensators in achieving practical implementations with a desired performance. Generally speaking, the inherent features of
a gravity compensator are affected by three main aspects: kinematic structure, motion/force transmissions, and passive force gen­
eration. Therefore, the design of a prospective gravity compensator should consider these three aspects to mitigate the limitations that
appear in the conventional approaches.
This paper proposes a design approach for gravity compensators using planar four-bar mechanisms and a linear spring. This work
constructs a class of gravity compensators with consideration of the following design criteria: (i) a simple kinematic structure is used
for the spring installation, (ii) only lower pairs are adopted for the motion/force transmissions, and (iii) a practical (non-ZFL) spring is
used for the passive force generation. With these criteria, the proposed gravity compensators can achieve good performance and
kinematic simplicity, providing them with the potential for practical implementation. More specifically, the construction of the gravity
compensators begins by combining planar four-bar mechanisms with a rotating mass. By attaching a spring to each mechanism and
then permuting the springs, 42 different types of gravity compensators are realized. The parameters of the gravity compensators are
determined via an optimization procedure for minimizing the actuator torque within a specified balancing zone. This study illustrates
the performances of the proposed gravity compensators via numerical examples and experiments, and their application for the gravity
compensation of serial robots.
The rest of this paper is organized as follows. Sections 2 and 3 present the construction and optimal design of the gravity com­
pensators, respectively. Then, the performance analysis of the gravity compensators via numerical examples is shown in Section 4.
Subsequently, Sections 5 and 6 respectively illustrate experiments with the gravity compensators and their application to serial robots.
Finally, the discussion and conclusion of this paper are drawn in Sections 7 and 8, respectively.

2. Construction of gravity compensators

Let us consider a rotating link with a mass to be gravity compensated. To attach a linear spring to the rotating link for gravity
compensation, an auxiliary mechanism is required to regulate the elastic torque generated by the spring to oppose the gravitational
torque caused by the link’s mass. This study adopts planar four-bar mechanisms for the passive force regulation as they possess a
simple kinematic structure. In each four-bar mechanism, the permutation of the spring (i.e., the change of the spring ends articulated to
the links) can acquire different spring designs, resulting in various gravity compensators. The construction process of the gravity
compensators is illustrated in Fig. 1 and is described as follows:

1 Step 1. Give a rotating link with a mass to be gravity compensated (Fig. 1(a)).

4
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 3. Free-body diagram of the type-E gravity compensator.

2 Step 2. Form planar four-bar mechanisms with lower pairs only (revolute (R) and prismatic (P) joints are taken). In this step, the
combination and permutation of four links and four R/P joints can generate ten different planar four-bar mechanisms, as shown in
Figs. 1(b-i)-(b-x).
3 Step 3. Combine a rotating link with a planar four-bar mechanism so that the pivot point of the rotating link coincides with an R
joint in the four-bar mechanism. Here, assume that link 1 is the fixed base and link 4 is connected to the rotating link. These as­
sumptions indicate that the connection joint between links 1 and 4 must be an R joint as a constraint for choosing a suitable four-bar
mechanism. Accordingly, one can realize that the RRRR, PRRR, RPRR, PPRR, PRPR, RPPR, and PPPR mechanisms (Figs. 1(b-i)-(b-
vii)) are satisfied with the above constraint. For instance, the combination of a rotating link with an RRRR mechanism is illustrated
in Fig. 1(c-iii). In this design, links 1, 2, 3, and 4 are called the fixed base, follower, coupler link, and crank (i.e., the input link whose
motion is induced by gravity), respectively.
4 Step 4. Attach a linear spring to each combined four-bar mechanism (in Step 3) and then permute the spring in the mechanism. A
spring is arbitrarily attached from one link to another in each four-bar mechanism. For example, a spring is connected from link 1 to
link 2 in a combined RRRR mechanism, as shown in Fig. 1(d). By permuting the spring in this mechanism, six different types of
gravity compensators (Group I) are obtained, as illustrated in Fig. 2. Similarly, six other groups, each having six different types, are
also derived from the other satisfied four-bar mechanisms (in Step 3).

In sum, the construction process described above can generate a class of gravity compensators encompassing 42 different types.

3. Optimal design for gravity compensation

The optimal design of a gravity compensator aims to seek the dimensions of the links and the spring parameters to minimize its
actuator torque during operations. This section takes types E and F in Group I (Figs. 2(e) and 2(f)) as two studied cases. Generally
speaking, the auxiliary four-bar mechanism of a gravity compensator should be smaller than the rotating (main) link in size and
weight. These requirements can improve the compactness of the gravity compensator while reducing the supplement weight attached
to the rotating link. In such a situation, a high-stiffness spring may be required if it is connected between two links in the four-bar
mechanism. The links and joints of the mechanism should be made with high rigidity to deal with the high elastic forces generated
by the spring. This need may increase the size and weight of the links and joints of the mechanism. A practical solution to mitigate the
above-described limitations is to articulate one spring end to the rotating link (which is adopted in the studied gravity compensators).
This arrangement can provide sufficient space for the spring to generate a desired elastic torque (or store enough potential energy) for
gravity compensation.

5
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

3.1. Static modeling

The free-body diagram method [51] is used to establish the static models of the type-E and type-F gravity compensators. In this
method, the equilibrium equation of forces for a body link i in a mechanism under static conditions can be written as:
∑ ∑
Fij + Fik + Fai + mi g = 0 (1)
j k

where Fij, Fik, and Fai represent the reaction force vector, the external force vector, and the actuator force vector exerted on link i,
respectively; mi and g the mass of link i and the gravitational acceleration vector, respectively. Note that Fik may include the gravi­
tational forces caused by other links and the elastic forces.
Next, the equilibrium equation of torques (or moments) for link i can be expressed as:
∑( ) ∑
rij × Fij + (rik × Fik ) + mi (rci × g) + Tai + Tri = 0 (2)
j k

where rij, rik, and rci stand for the position vectors associated with the reaction force Fij, the external force Fik, and the mass center of
link i, respectively; Tai and Tri the actuator torque vector and reaction torque vector exerted on link i, respectively. Note that the
equilibrium equation of moments (Eq. (2)) can be applied to any origin on a link.

3.1.1. Type-E gravity compensator


The free-body diagram of the type-E gravity compensator is shown in Fig. 3. In this design, a coordinate system (x, y) is placed at
the pivot point O of the rotating link. A spring is connected from link 2 (at point D) to the rotating link (at point H). The rotating link is
assumed to operate within a range of joint angles θ = (θin, θen), which is defined as the balancing zone of the gravity compensator.
According to Hooke’s law, the elastic force Fs generated at point D can be expressed as:
Fs = k(s0 − d0 + d1 ) (3)

where
[ ] [ ]
xD − l1 + l2a cos(α + ε)
d0 = DH|θin =‖ vH − vD ‖θin , d1 = DH|(θin , θen ) =‖ vH − vD ‖(θin , θen ) , vD = = ,
yD l2a sin(α + ε)
[ ] [ ] (2 ) (4)
xH ls cosθ 1 l2 + l22a − l22b
vH = = , ε = cos− , l1 = OA, l2 = AB, l2a = AD, l2b = BD, ls = OH
yH ls sinθ 2l2 l2a

In Eqs. (3) and (4), k, s0, d0, and d1 stand for the spring stiffness coefficient, the initial spring extension, the distances DH at the
initial and instantaneous positions of the rotating link, respectively; vD and vH the position vectors of points D and H, respectively; α the
joint angle of link 2, as calculated in Appendix A.
From Eq. (3), the elastic force vector Fs along the longitudinal axis of the spring is written as:
[ ] [ ]
Fsx Fs cosγd
Fs = = (5)
Fsy Fs sinγd

where
( ) [ ]
x ⋅ μs vH − vD 1
γd = cos− 1
, μs = , x= (6)
‖ x ‖‖ μs ‖ ‖ vH − vD ‖ 0

Now, the equilibrium equations of forces and moments Eqs. (1) and ((2)) are used to determine the forces and torques exerted on
the mechanism. First, the equilibrium of forces on link 2 can be expressed as:
− F1 − F3 + Fs + m2 g = 0 (7)

where F1 and F3 represent the elastic force vectors applied to link 1 at joint A and link 3 at joint B, respectively. Note that –F1 and –F3
are the reaction forces on link 2 at joints A and B, respectively. Eq. (7) can be transformed into two algebraic equations, i.e.:
F1x + F3x − Fsx = 0 (8)

F1y + F3y − Fsy + m2 g = 0 (9)

Similarly, the equilibrium of forces on link 3 can be written as:


F3 − F4 + m3 g = 0 (10)

where F4 represents the elastic force vector applied to link 4 at joint C. Then, two algebraic equations can be derived from Eq. (10), i.e.:

6
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

F3x − F4x = 0 (11)

F3y − F4y − m3 g = 0 (12)

Next, the equilibrium of moments (Eq. (2)) on link 2 at joint A can be expressed as:
(vD − vA ) × Fs + (vB − vA ) × (− F3 ) + (vP2 − vA ) × (m2 g) + Ta2 = 0 (13)

where
[ ] [ ] [ ] [ ] [ ]
xA − l1 x − l1 + l2 cosα x vA + vB + vD
vA = = , vB = B = , vP2 = P2 = (14)
yA 0 yB l2 sinα yP2 3
In Eqs. (13) and (14), vA, vB, and vP2 stand for the position vectors of points A, B, and P2 (i.e., the mass center of link 2), respectively.
Eq. (13) is rewritten as:
[ ] [ ] [ ] [ ] [ ] [ ]
xD − xA Fsx xB − xA F3x x − xA 0
× − × + P2 × + Ta2 = 0 (15)
yD − yA Fsy yB − yA F3y yP2 − yA − m2 g

or equivalently
(xD − xA )Fsy − (yD − yA )Fsx − (xB − xA )F3y + (yB − yA )F3x − (xP2 − xA )m2 g + Ta2 = 0 (16)

Similarly, the equilibrium of moments (Eq. (2)) on link 2 at joint B is given as:
(vD − vB ) × Fs + (vA − vB ) × (− F1 ) + (vP2 − vB ) × (m2 g) + Tr2 = 0 (17)
Eq. (17) is rewritten as:
[ ] [ ] [ ] [ ] [ ] [ ]
xD − xB Fsx xA − xB F1x x − xB 0
× − × + P2 × + Tr2 = 0 (18)
yD − yB Fsy yA − yB F1y yP2 − yB − m2 g

or equivalently
(xD − xB )Fsy − (yD − yB )Fsx − (xA − xB )F1y + (yA − yB )F1x − (xP2 − xB )m2 g + Tr2 = 0 (19)

The equilibrium of moments (Eq. (2)) on link 3 at joint B is expressed as:


(vC − vB ) × (− F4 ) + (vP3 − vB ) × (m3 g) − Tr2 = 0 (20)

where
[ ] [ ] [ ]
xC l4 cos(θ + σ ) x vB + vC
vC = = , vP3 = P3 = (21)
yC l4 sin(θ + σ) yP3 2
In Eq. (21), vC and vP3 stand for the position vectors of points C and P3 (i.e., the mass center of link 3), respectively. Eq. (20) is
rewritten as:
[ ] [ ] [ ] [ ]
xC − xB F4x x − xB 0
− × + P3 × − Tr2 = 0 (22)
yC − yB F4y yP3 − yB − m3 g

or equivalently
− (xC − xB )F4y + (yC − yB )F4x − (xP3 − xB )m3 g − Tr2 = 0 (23)

Assume that joint O is an active joint, where a motor is placed to actuate the gravity compensator. This assumption suggests that the
actuator torque at joint A should be zero (Ta2 = 0). Let ΩX = [F1x, F1y, F3x, F3y, F4x, F4y, Tr2]T denote a wrench vector to express the
forces and moments exerted on the gravity compensator. By arranging Eqs. (8), (9), (11), (12), (16), (19), and (23) into a matrix form, a
linear matrix equation is obtained as:
ΩA ⋅ ΩX = ΩB (24)

where
⎡ ⎤
1 0 1 0 0 0 0
⎢ 0 1 0 1 0 0 0 ⎥
⎢ ⎥
⎢ 0 0 1 0 − 1 0 0 ⎥
⎢ ⎥
ΩA = ⎢
⎢ 0 0 0 1 0 − 1 0 ⎥⎥ (25)
⎢ 0 0 yB − yA xA − xB 0 0 0 ⎥
⎢ ⎥
⎣ yA − yB xB − xA 0 0 0 0 1 ⎦
0 0 0 0 yC − yB xB − xC − 1

7
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 4. Free-body diagram of the type-F gravity compensator.

⎡ ⎤
Fsx
⎢ Fsy − m2 g ⎥
⎢ ⎥
⎢ 0 ⎥
⎢ ⎥

ΩB = ⎢ m3 g ⎥
⎥ (26)
⎢ (yD − yA )Fsx − (xD − xA )Fsy + (xP2 − xA )m2 g ⎥
⎢ ⎥
⎣ (yD − yB )Fsx − (xD − xB )Fsy + (xP2 − xB )m2 g ⎦
(xP3 − xB )m3 g

When the matrix ΩA is nonsingular and invertible, the vector ΩX can be found from Eq. (24) as:

ΩX = Ω−A 1 ⋅ ΩB (27)

3.1.2. Type-F gravity compensator


Fig. 4 depicts the free-body diagram of the type-F gravity compensator. In this design, a spring is connected from link 3 (at point D)
to the rotating link (at point H). Similar to Eq. (5), the elastic force vector Fs along the longitudinal axis of the spring can be formulated,
but it must consider a new position vector of point D, i.e.:
[ ] [ ]
x − l1 + l2 cosα + l3a cos(β + ε)
vD = D = (28)
yD l2 sinα + l3a sin(β + ε)

where
(2 )
l3 + l23a − l23b
ε = cos− 1
, l3 = BC, l3a = BD, l3b = CD (29)
2l3 l3a

and new mass centers of links 2 and 3, i.e.:


[ ] [ ]
x vA + vB x vB + vC + vD
vP2 = P2 = , vP3 = P3 = (30)
yP2 2 yP3 3
Similar to Section 3.1.1, the forces and torques exerted on type F can be determined. First, the equilibrium of forces (Eq. (1)) on
links 2 and 3 can respectively be expressed as:
− F1 + m2 g + F2 = 0 (31)

8
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

− F2 − F4 + Fs + m3 g = 0 (32)

where F1, F2, and F4 represent the elastic force vectors applied to link 1 at joint A, link 2 at joint B, and link 4 at joint C, respectively.
Then, the equilibrium of moments (Eq. (2)) on links 2 at joints A and B and link 3 at joint B can respectively be written as:
(vB − vA ) × F2 + (vP2 − vA ) × (m2 g) + Ta2 = 0 (33)

(vA − vB ) × (− F1 ) + (vP2 − vB ) × (m2 g) − Tr3 = 0 (34)

(vD − vB ) × Fs + (vC − vB ) × (− F4 ) + (vP3 − vB ) × (m3 g) + Tr3 = 0 (35)


Let ΨX = [F1x, F1y, F2x, F2y, F4x, F4y, Tr3]T denote a wrench vector to express the forces and moments exerted on the gravity
compensator. Eqs. (31)–(35) can be transformed into seven algebraic equations, as detailed in Appendix B. When the actuator torque at
joint A is assumed to be zero (Ta2 = 0), these algebraic equations can be arranged into a linear matrix equation as:
ΨA ⋅ ΨX = ΨB (36)

where
⎡ ⎤
− 1 0 1 0 0 0 0
⎢ 0 − 1 0 1 0 0 0 ⎥
⎢ ⎥
⎢ 0 0 1 0 1 0 0 ⎥
⎢ ⎥
ΨA = ⎢
⎢ 0 0 0 1 0 1 0 ⎥
⎥ (37)
⎢ 0 0 yA − yB xB − xA 0 0 0 ⎥
⎢ ⎥
⎣ yA − yB xB − xA 0 0 0 0 − 1⎦
0 0 0 0 yC − yB xB − xC 1
⎡ ⎤
0
⎢ m2 g ⎥
⎢ ⎥
⎢ Fsx ⎥
⎢ ⎥
ΨB = ⎢
⎢ F sy − m3 g ⎥
⎥ (38)
⎢ (xP2 − xA )m2 g ⎥
⎢ ⎥
⎣ (xP2 − xB )m2 g ⎦
(yD − yB )Fsx − (xD − xB )Fsy + (xP3 − xB )m3 g

From Eq. (36), the vector ΨX can be written as:

Ψ X = Ψ A− 1 ⋅ Ψ B (39)

when the matrix ΨA is nonsingular and invertible.

3.1.3. Actuator torques


When the wrench vectors ΩX (Eq. (27) for type E) and ΨX (Eq. (39) for type F) are known, the actuator torques of the gravity
compensators can be determined. According to Eq. (2), the equilibrium of moments on link 4 at joint O can be expressed as:
vC × F4 + vP 4 × (m4 g) + vH × (− Fs ) + vG × (mc g) + Ta4 = 0 (40)

where
[ ] [ ] [ ]
xP4 vC + vE x l cosθ
vP4 = = , vG = G = c , lc = OG (41)
yP4 3 yG lc sinθ

In Eq. (41), vP4 and vG represent the position vectors of points P4 (i.e., the mass center of link 4) and G (i.e., the mass center of the
rotating link), respectively.
From Eq. (40), the actuator torque Ta4 at joint O can be written as:
[ ] [ ] [ ] [ ] [ ] [ ] [ ] [ ]
xC F4x xP4 0 xH Fsx xG 0
Ta4 = − × − × + × − ×
yC F4y yP4 − m4 g yH Fsy yG − mc g (42)
= − xC F4y + yC F4x + xP4 m4 g + xH Fsy − yH Fsx + xG mc g

Eq. (42) is valid for both types E and F. The actuator torque Ta4 can be called the compensated torque as it takes the elastic force Fs
(Eq. (5)) into account.
Consequently, the uncompensated torque Tun is defined as the actuator torque at joint O without consideration of the elastic force
Fs. From Eq. (42), the uncompensated torque Tun is expressed as:
Tun = − vG × (mc g) = xG mc g (43)

9
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 5. Spring displacement at different positions of the rotating link.

3.2. Optimization problem

In this study, the link lengths and spring parameters of the type-E and type-F gravity compensators are considered their design
variables, which are expressed as:

X = [l1 , l2 , l3 , l4 , ls , l2a , l2b , l4a , l4b , k, s0 ]T for type E (44)

Y = [l1 , l2 , l3 , l4 , ls , l3a , l3b , l4a , l4b , k, s0 ]T for type F (45)


For brevity, this section only presents the optimization problem of type E, while that of type F can be done similarly.

3.2.1. Objective function


Let δt denote the ratio of the maximum compensated torque to the maximum uncompensated torque, in absolute value, of the
gravity compensator within a balancing zone (θin, θen). From Eqs. (42) and (43), the ratio δt is formulated as:

max{|Ta4 |}⃒⃒
δt = (46)
max{|Tun |}⃒(θin , θen )

In Eq. (46), δt is called the torque reduction ratio (TRR), which is inversely proportional to the performance of gravity compen­
sation. The value δt = 0 means that the gravitational torque is completely eliminated by the elastic torque at all the positions in the
balancing zone (i.e., complete gravity compensation). The gravity compensator deteriorates its performance when δt > 1. Therefore,
the objective of the optimization problem is to minimize the TRR δt.

3.2.2. Constraints
To form a single-loop mechanism with four links (Fig. 3), the length of a link must be less than the total length of the other three
links, i.e.:


⎪ l1 − l2 − l3 − l4 < 0

− l1 + l2 − l3 − l4 < 0
(47)

⎪ − l − l2 + l3 − l4 < 0
⎩ 1
− l1 − l2 − l3 + l4 < 0

For the triangle links 2 and 4, the length of an edge must be less than the total length of the other two edges, i.e.:


⎪ l2 − l2a − l2b < 0



⎪ − l2 + l2a − l2b < 0

− l2 − l2a + l2b < 0
(48)

⎪ l4 − l4a − l4b < 0



⎪ − l4 + l4a − l4b < 0

− l4 − l4a + l4b < 0

As shown in Fig. 3, the spring in type E should be attached to the rotating link. Then, a constraint for the spring end at point H can

10
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Table 1
Prescribed parameters of types E and F with balancing zone [0, π/2].
mc (kg) m2 (kg) m3 (kg) m4 (kg) lc (m) ρs (-) θ (rad)

0.5 0.05 0.05 0.05 0.2 0.35 [0, π/2]

Table 2
Design variables and optimization results of the gravity compensators with balancing zone [0, π/2].
Parameters Unit Bounds Optimal values
Lower Upper Type E Type F

δt – – – 0.015 0.02
l1 m 0.02 0.06 0.0331 0.0593
l2 m 0.02 0.06 0.0348 0.0495
l3 m 0.02 0.06 0.0554 0.0570
l4 m 0.02 0.06 0.0219 0.0463
ls m 0.06 0.18 0.1790 0.1736
l2a m 0 0.06 0.06 –
l2b m 0 0.06 0.0261 –
l3a m 0 0.06 – 0.0569
l3b m 0 0.06 – 0.0287
l4a m 0 0.06 0.0186 0.0198
l4b m 0 0.06 0.0033 0.0483
k N/m 100 2000 740.7 817.5
s0 m 0 0.06 0 0.0052

Fig. 6. Optimization results of the gravity compensators with balancing zone [0, π/2]: (a) convergence of the best objectives and (b) actu­
ator torques.

yield as:
− ls + l4a < 0 (49)
From Eqs. (47)–(49), the link constraints of type E can be arranged into a matrix form as:
Φ⋅X< 0 (50)

where

11
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 7. Optimal configurations of the gravity compensators with balancing zone [0, π/2]: (a) type E and (b) type F.

Fig. 8. Free lengths and extensions of the springs in the gravity compensators with balancing zone [0, π/2]:
(a) type E and (b) type F.

⎡ ⎤
1 − 1 − 1 − 1 0 0 0 0 0 0 0
⎢− 1 1 − 1 − 1 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢− 1 − 1 1 − 1 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢− 1 − 1 − 1 1 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ 0 1 0 0 0 − 1 − 1 0 0 0 0⎥
⎢ ⎥
Φ=⎢
⎢ 0 − 1 0 0 0 1 − 1 0 0 0 0⎥⎥ (51)
⎢ 0 − 1 0 0 0 − 1 1 0 0 0 0⎥
⎢ ⎥
⎢ 0 0 0 1 0 0 0 − 1 − 1 0 0⎥
⎢ ⎥
⎢ 0 0 0 − 1 0 0 0 1 − 1 0 0⎥
⎢ ⎥
⎣ 0 0 0 − 1 0 0 0 − 1 1 0 0⎦
0 0 0 0 − 1 0 0 1 0 0 0
To have a feasible choice for the spring in the gravity compensator, a spring constraint is established as follows. The spring
displacement at different positions of the rotating link within a balancing zone (θin, θen) is illustrated in Fig. 5. Here, the free length L0 of
the spring is calculated as:

12
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 9. MSC Adams models of the gravity compensators: (a) type E and (b) type F.

Fig. 10. MSC Adams simulated actuator torques of the gravity compensators: (a) type E and (b) type F.

L0 = d0 − s0 (52)

where d0 and s0 stand for the spring length at the initial position θin and the initial spring extension, respectively.
Generally, any commercially available spring has a limit to describe how far it can be extended. This study assumes that the
maximum spring extension within a balancing zone is defined by an extension-to-length ratio (ETLR) ρs, which is expressed as:
max{d1 } − L0
≤ ρs , ρs ≥ 0 (53)
L0
Eq. (53) implies that the ETLR is proportional to the extension capability of the spring. For instance, when ρs = 0 is defined, the
spring will not be extended within the balancing zone. The spring extension can equal the initial length L0 when ρs = 1 is set.
With the substitution of Eq. (52) into Eq. (53), a constraint for the spring extension is obtained as:
max{d1 } − (ρs + 1)(d0 − s0 ) ≤ 0 (54)

3.2.3. Problem formulation


From the design variable vector Eq. (44)), the objective function (Eq. (46)), the link and spring constraints (Eqs. (50) and ((54)), the
optimization problem of type E is formulated as:

13
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 11. Torque reduction ratios (TRRs) of the gravity compensators with different balancing zones:
(a) type E and (b) type F.

Table 3
Optimization results of type E with different balancing zones.
Parameters Unit Zone 1 Zone 2 Zone 3 Zone 4 Zone 5 Zone 6

θ rad [0, π/2] [–π/2, 0] [–π/4, π/4] [–π/4, π/2] [–π/2, π/4] [–π/2, π/2]
δt – 0.015 0.1299 0.0698 0.2123 0.1935 0.2497
l1 m 0.0331 0.0269 0.0553 0.0365 0.0297 0.0251
l2 m 0.0348 0.06 0.0467 0.0362 0.0505 0.0371
l3 m 0.0554 0.0257 0.0373 0.0588 0.0586 0.0482
l4 m 0.0219 0.02 0.0218 0.0269 0.02 0.022
ls m 0.179 0.18 0.18 0.18 0.18 0.18
l2a m 0.06 0.0599 0.06 0.06 0.0581 0.0561
l2b m 0.0261 0.0040 0.0397 0.0317 0.0142 0.0269
l4a m 0.0186 0.0497 0.0349 0.0418 0.0519 0.0267
l4b m 0.0033 0.0499 0.0134 0.015 0.032 0.005
k N/m 740.7 817.5 817.5 713.77 794.49 817.5
s0 m 0 0.0217 0.0169 0 0.0167 0

Table 4
Optimization results of type F with different balancing zones.
Parameters Unit Zone 1 Zone 2 Zone 3 Zone 4 Zone 5 Zone 6

θ rad [0, π/2] [–π/2, 0] [–π/4, π/4] [–π/4, π/2] [–π/2, π/4] [–π/2, π/2]
δt – 0.02 0.022 0.058 0.174 0.136 0.327
l1 m 0.0593 0.0312 0.0304 0.0244 0.037 0.048
l2 m 0.0495 0.0494 0.0579 0.0557 0.049 0.0401
l3 m 0.0570 0.0431 0.0394 0.0535 0.0575 0.0555
l4 m 0.0463 0.032 0.0484 0.0293 0.0277 0.0251
ls m 0.1736 0.178 0.1799 0.18 0.18 0.18
l3a m 0.0569 0.0482 0.0458 0.0349 0.0507 0.0499
l3b m 0.0287 0.0442 0.0363 0.058 0.0427 0.0264
l4a m 0.0198 0.0103 0.0163 0.0386 0.018 0.0319
l4b m 0.0483 0.0409 0.0591 0.0399 0.0403 0.0397
k N/m 817.5 591.58 743.23 803.72 812.69 817.5
s0 m 0.0052 0.0282 0.0195 0 0.0123 0

14
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 12. Optimal configurations of type E with different balancing zones: (a) zone 1, (b) zone 2, (c) zone 3, (d) zone 4, (e) zone 5, and (f) zone 6.

Minimize δt (X)|θin ≤ θ ≤ θen


with X = [l1 , l2 , l3 , l4 , ls , l2a , l2b , l4a , l4b , k, s0 ]T ∈ R
subject to Φ ⋅ X < 0 (55)
max{d1 } − (ρs + 1)(d0 − s0 ) ≤ 0
Xj , d0 > 0; s0 , ρs ≥ 0; ζlj ≤ Xj ≤ ζuj (j = 1, 2, ..., 11)

where ζlj and ζuj stand for the lower and upper bounds of component Xj (j = 1, 2,…, 11) in the design variable vector X (Eq. (44)).

4. Performance analysis

This section analyzes the performances of the type-E and type-F gravity compensators via numerical examples. The variation of the
balancing zone and the use of different optimization algorithms are also discussed.

4.1. Design for a specific balancing zone

Let us consider the types-E and type-F gravity compensators whose prescribed parameters are presented in Table 1. The bounds of
their design variables are listed in Table 2. This study used a Genetic algorithm (GA) in MATLAB Toolbox to solve the optimization
problem in Eq. (55). Ten independent runs were executed for each selected type, and the one with the best objective value was taken.
The optimization results are also presented in Table 2.
Fig. 6(a) illustrates the convergence of the best objectives of types E and F. The best objective values of types E and F can achieve up

15
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 13. Optimal configurations of type F with different balancing zones: (a) zone 1, (b) zone 2, (c) zone 3, (d) zone 4, (e) zone 5, and (f) zone 6.

Table 5
Optimization results of type F using the PSO and DE algorithms.
Alg. δt (-) l1 (m) l2 (m) l3 (m) l4 (m) ls (m) l3a (m) l3b (m) l4a (m) l4b (m) k (N/m) s0 (m)

PSO 0.01 0.0218 0.0489 0.0374 0.0288 0.1621 0.0298 0.0292 0.0262 0.0292 1452.9 0
DE 0.05 0.026 0.06 0.0367 0.04 0.18 0.034 0.0264 0.0227 0.0279 1443 0.014

to 0.015 and 0.02, respectively. With the optimal parameters listed in Table 2, both types E and F significantly reduce their actuator
torques, as shown in Fig. 6(b). The optimal configurations of these two types are plotted in Fig. 7.
Fig. 8 shows the free lengths and extensions of the springs in types E and F. In type E (Fig. 8(a)), the maximum spring extension is
0.052 m, and the free length of the spring is 0.159 m. By adopting Eq. (53), the calculated ETLR of the spring is given as 0.327, which is
lower than the prescribed ETLR (ρs = 0.35). Similarly, the computed ETLR of the spring in type F (Fig. 8(b)) is also lower than the
prescribed ETLR. The obtained results show that the optimal parameters of types E and F (Table 2) are satisfied with the spring
constraint (Eq. (54)).
Next, MSC Adams software was used to verify the performances of the gravity compensators whose parameters are listed in Tables 1
and 2. The simulation models of types E and F are shown in Figs. 9(a) and 9(b), respectively. Assume that the gravity compensators
moved from the position θ = π/2 (initial position) to θ = 0 (end position) within 9 s. The simulated actuator torques of types E and F are
presented in Figs. 10(a) and 10(b), respectively. The results show that the simulated torque curves of types E and F are similar to their
calculated torque curves (Fig. 6(b)). The simulated TRRs can yield δt = 0.01 by type E and δt = 0.02 by type F.

16
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 14. Convergence of the best objective of type F using different optimization algorithms.

4.2. Variation of the balancing zone

This section discusses the performances of the gravity compensators with different balancing zones. Here, six balancing zones are
studied: [0, π/2] as zone 1, [–π/2, 0] as zone 2, [–π/4, π /4] as zone 3, [–π/4, π/2] as zone 4, [–π/2, π/4] as zone 5, and [–π/2, π /2] as
zone 6. Similar to Section 4.1, a GA algorithm was used to solve the optimization problem in Eq. (55), and ten independent runs were
also executed for each selected balancing zone. Fig. 11 shows the TRRs of types E and F with different balancing zones. Generally, the
TRR is proportional to the size of the balancing zone, i.e., the performance of gravity compensation is reduced when enlarging the
balancing zone. Zone 1 provides the highest performance in which the TRR can achieve up to 0.015 by type E and 0.02 by type F. In
contrast, zone 6 exhibits the lowest performance with a mean TRR of 0.35 by type E and 0.36 by type F. The optimization results of
types E and F with different balancing zones are detailed in Tables 3 and 4, respectively. Their optimal configurations are plotted in
Figs. 12 and 13.

4.3. Use of different optimization algorithms

This section exploits the Particle swarm optimization (PSO) [52,53] and Differential evolution (DE) [54,55] algorithms to solve the
optimization problem in Eq. (55). The PSO and DE algorithms are described in Appendix C and Appendix D, respectively. This study
only considers type F (Fig. 4) with balancing zone [0, π/2]. A population size of 500 was set for the PSO and DE algorithms. Each
algorithm was executed with ten independent runs, and the one with the best objective value was selected. The optimization results of
type F using the PSO and DE algorithms are presented in Table 5. The convergence of the best objective and the actuator torque are
illustrated in Figs. 14 and 15, respectively. The results show that the PSO provides the best objective value (δt = 0.01), but it requires
1000 iterations to achieve this value. Within the first 200 iterations, the PSO appears the fastest convergence, whereas the DE has the
slowest one. The GA produces a more consistent result of the actuator torque as compared with the two others.

5. Experiments

Prototypes of the type-E and type-F gravity compensators were built (using a 3D printer) to demonstrate their design feasibilities
and performances. Similar to Section 4, the prescribed parameters and optimization results of the gravity compensators are presented
in Tables 6 and 7, respectively. Experiments with the gravity compensators were first executed via passive tests in which types E and F
were manually operated within balancing zone [–π/4, π /2], as illustrated in Fig. 16. The results show that both types E and F could
maintain their configurations at specific positions θ = {π /2, π/3, π /6, 0, –π/6, –π/4}. A video demonstrating the passive tests is
provided in the supplementary material of this paper.
Fig. 17 shows an experimental setup for an active test with type E. In this setup, a DC motor (MAXON EC-i40) integrated with a

17
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 15. Actuator torque of type F using different optimization algorithms: (a) GA, (b) PSO, and (c) DE.

Table 6
Prescribed parameters of the prototypes of types E and F.
mc (kg) m2 (kg) m3 (kg) m4 (kg) lc (m) θ (rad)

0.2 0.01 0.01 0.01 0.2 [–π/4, π/2]

18
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Table 7
Optimization results of the prototypes of the gravity compensators.
Parameters Unit Bounds Optimal values
Lower Upper Type E Type F

δt – – – 0.0185 0.0215
l1 m 0.02 0.06 0.0209 0.0222
l2 m 0.02 0.06 0.0388 0.0510
l3 m 0.02 0.06 0.0469 0.0488
l4 m 0.02 0.06 0.0253 0.0265
ls m 0.06 0.18 0.1592 0.1571
l2a m 0.02 0.06 0.0559 –
l2b m 0.02 0.06 0.0272 –
l3a m 0.02 0.06 – 0.0268
l3b m 0.02 0.06 – 0.0347
l4a m 0.02 0.06 0.0588 0.0321
l4b m 0.02 0.06 0.0339 0.0233
k N/m – – 790 790
s0 m – – 0 0

Fig. 16. Passive tests with the gravity compensators: (a) type E and (b) type F.

gearbox (MAXON GP42C) and an encoder (MAXON 16EASY) was used to actuate the gravity compensator. A positioning controller
(MAXON EPOS4 Module 50/15) was used to control the motor. The gravity compensator was controlled to move from the position θ =
π/2 to θ = –π/4 within 6 s and then turned back within the same time. An active test with type F was done similarly. During the
operation of each gravity compensator, the actuator (input) torque of the motor was acquired from EPOS Studio software on a
computer. The actuator torques of types E and F are shown in Fig. 18. One can see that the actuator torques of these two types are
significantly reduced throughout their operations. The peak uncompensated and compensated torques of type E are 0.51 and 0.069 N-
m, leading to a torque reduction rate of 86.5%. The maximum compensated torque of type F is 0.062 N-m, which indicates a torque
reduction rate of 87.8%.

6. An application: gravity compensation of serial robots

This section presents an application of the proposed gravity compensators to serial robots. More specifically, type F (Fig. 4) is
adopted for the gravity compensation design for a KUKA KR 210 R3100 robot [56]. The performance of the robot is evaluated to show
the effectiveness of the proposed design. A prototype of a serial robot is also presented to validate the applicability of the gravity
compensators.

6.1. Design concept

As presented in [57], the gravitational torques at the shoulder (2nd) and elbow (3rd) joints of the KUKA robot are much more
significant than those at the other joints. Consequently, the proposed gravity compensation design for the robot only considers the

19
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 17. Experimental setup for an active test with type E.

Fig. 18. Measured actuator torques of types E and F.

shoulder and elbow joints, as shown in Fig. 19. In this design, a type-F gravity compensator i (i = 1, 2) is connected to link i + 1 so that
joint Oi is coincident with the shoulder (i = 1) or elbow (i = 2) joint. Link OiEi aligns with the longitudinal axis of link i + 1. An auxiliary
link A1P2 is added to form a parallelogram linkage (O1O2P2A1) to maintain the reference frame of the second compensator (i.e., link
O2A2 in the horizontal direction). A spring ki is connected from link BiCi at point Di to link i + 1 at point Hi.
By taking advantage of the parallelogram O1O2P2A1, the gravity compensation design for each (shoulder/elbow) joint can be
realized by using a single gravity compensator. This design strategy has widely been applied to serial robots [19,28,30,40]. In the
present work, the accumulated rotating masses, whose centers are located on links i + 1 (i = 1, 2), should be identified first. Let MCi and
LCi denote the accumulated rotating mass applying to joint Oi and the distance from the accumulated mass center to joint Oi,
respectively. From the distribution of mass in space, the parameters MCi and LCi are expressed as:

3
MCi = mci (56)
i

[ ( ) ]
1 ∑3
LCi = mci lci + mcj Li (57)
MCi i+1

where mci, Li, and lci (i = 1, 2) represent the mass, the length of link i + 1, and the distance from the mass center Gi of link i + 1 to joint
Oi; mc3 the mass of the payload. The dimensional parameters and masses of the robot were estimated from its CAD model, as listed in
Table 8.

20
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 19. Gravity compensation design of the KUKA robot using the type-F gravity compensators.

Table 8
Dimensional parameters and masses (estimated) of the KUKA robot.
L1 (m) L2 (m) lc1 (m) lc2 (m) mc1 (kg) mc2 (kg) mc3 (kg)

1.35 1.4 0.525 0.482 349 324 30

Similar to Section 3, the compensated torque Ta4i and the uncompensated torque Tuni at joint Oi (i = 1, 2) can be written as:

2
[ ]
Ta4i = − (xCi − xOi )F4yi + (yCi − yOi )F4xi + (xP4i − xOi )m4i g + (xHi − xOi )Fsyi − (yHi − yOi )Fsxi + LCi Mi gcosθi (58)
i


2
Tuni = LCi Mi gcosθi (59)
i

where vCi = [xCi, yCi]T, vOi = [xOi, yOi]T, vP4i = [xP4i, yP4i]T, and vHi = [xHi, yHi]T represent the position vectors of points Ci, Oi, P4i, and
Hi, respectively; F4i = [F4xi, F4yi]T and Fsi = [Fsxi, Fsyi]T the elastic force vectors applied to link 4i at joint Ci and along the longitudinal
axis of the spring ki, respectively; θi the joint angle at joint Oi.
From Eqs. (58) and (59), the TRR δti at joint Oi can be determined similar to Eq. (46). The optimization problems of the two gravity
compensators attached to the robot are formulated like Eq. (55). The GA algorithm was also used to solve the optimization problems.
The results are presented in Table 9.

21
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Table 9
Optimization results of the gravity compensators attached to the KUKA robot.
Parameters Unit Joints Oi
Shoulder (i = 1) Elbow (i = 2)

δti – 0.0185 0.0184


θi rad [0, π/2] [–π/3, π/3]
m1i kg 15 10
m2i kg 15 10
m3i kg 15 10
m4i kg 15 10
m5i* kg 25 –
l1i m 0.1004 0.1323
l2i m 0.2504 0.2447
l3i m 0.2936 0.2371
l4i m 0.2837 0.2360
lsi m 1.1467 0.8604
l3ai m 0.2978 0.2489
l3bi m 0.1970 0.1295
l4ai m 0.2442 0.2804
l4bi m 0.1699 0.1457
ki N/mm 616.03 231.83
s0i m 0.1133 0.0653

* The mass of link A1P2.

Fig. 20. Total gravitational torques of the KUKA robot: (a) uncompensated and (b) compensated.

6.2. Performance evaluation

Fig. 20 illustrates the total gravitational torques of the robot over the studied workspace in the cases without and with gravity
compensation. A significant reduction in the gravitational torque of the robot, in absolute value, is achieved when the proposed gravity
compensation design is applied. The peak uncompensated and compensated torques, in absolute value, are 9096.4 and 167.8 N-m,
respectively. These results indicate that the actuator torque of the robot is reduced by 98.2%.
MSC Adams software was taken to verify the performance of the gravity-compensated robot. In this study, the robot was assumed to
track along two different trajectories, i.e., trajectory I in the pitch plane of the robot (Fig. 21(a)) and trajectory II in the three-
dimensional (3D) space (Fig. 21(b)). The simulated actuator torques of the robot during trajectory tracking are illustrated in
Fig. 22. One can see that the actuator torques, in absolute value, at joints O1 and O2 are considerably decreased with gravity
compensation. Along trajectory I, the peak uncompensated and compensated torques at joint O1 are 7232 and 164 N-m, respectively,
indicating a torque reduction rate of 97.7%. The peak actuator torque, in absolute value, at joint O2 is reduced from 2034 to 208 N-m,

22
V.L. Nguyen
23

Mechanism and Machine Theory 172 (2022) 104770


Fig. 21. MSC Adams simulation with the KUKA robot: (a) trajectory I and (b) trajectory II.
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 22. MSC Adams simulated actuator torques of the KUKA robot: (a) trajectory I and (b) trajectory II.

Fig. 23. Prototype and performance demonstration of a 2-DoF gravity-compensated serial robot.

showing a torque reduction rate of 89.8%. The actuator torques of the robot along trajectory II are also significantly reduced in which
the torque reduction rates at joints O1 and O2 are 89.3% and 95.9%, respectively.

6.3. A prototype

As presented in Section 6.1, the gravity compensation design for the KUKA robot considers only the second and third joints, whose
axes are normal to the pitch plane of the robot. Consequently, a 2-DoF serial robot was prototyped (using a 3D printer) to demonstrate
the applicability of the proposed gravity compensators, as shown in Fig. 23. The robot was assumed to carry a payload of 0.2 kg within

24
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Fig. 24. MSC Adams simulated actuator torques of the KUKA robot with different payloads:
(a) joint O1 and (b) joint O2.

a balancing zone defined by θ1 = [0, π/2] and θ2 = [–π/3, π /2]. A passive test with the prototype was then performed. It showed that
the robot could maintain its configuration (without input torques) at some positions in the prescribed balancing zone, as detailed in
Fig. 23. This result indicates the feasibility and good performance of the gravity compensation design.

7. Discussion: gravity compensation of robots with different payloads

Gravity compensation of robots with different payloads is significant to enhance the flexibility and adaptability of robots for various
applications [58–60]. As presented in Section 6, the proposed gravity compensation design for the KUKA robot was realized for a
constant (fixed) payload (30 kg). A question arising from this work is whether or not the proposed design is still effective when the
payload is changed. Now, let us consider different payloads applied to the robot: mc3 = {0, 15, 30, 45, 60, 75, 90} kg, which are
equivalent to {–100%, –50%, +0%, +50%, +100%, +150%, +200%} of the original payload (30 kg). To analyze the performance of
the robot with different payloads, MSC Adams simulation was performed in which the robot was assumed to track along trajectory II
(Fig. 21(b)). The actuator torques of the robot with different payloads are presented in Fig. 24. The compensated torques at joints O1
and O2 fluctuate from their original torques (i.e., the red curves) when the mass of the payload is added or subtracted from the original
payload (30 kg). The more difference between the applied payload and the original one, the lower the performance of gravity
compensation. The actuator torques and torque reduction rates of the robot with different payloads are detailed in Table 10. The lowest
compensated torques at joints O1 and O2 are obtained with the original payload (30 kg). This result is because the gravity compensation
design for the robot was realized for this payload, as presented in Section 6. The torque reduction rates at joints O1 and O2 with the
original payload are 89.3% and 95.9%, respectively. When the payload is detached (mc3 = 0 kg), the torque reduction rates reduce to
84.7% at joint O1 and 77.7% at joint O2. The torque reduction rates at joints O1 and O2 also decrease to 76.5% and 50.5%, respectively,
when the payload increases by 200% (mc3 = 90 kg). The obtained results imply that the proposed gravity compensation design for the
robot is still effective for payloads ranging from –100% to +200% of the original payload.

8. Conclusion

This paper proposed a design approach for gravity compensators using planar four-bar mechanisms and a linear spring. This work
rendered a class of gravity compensators with 42 different types, which possess high performance and a simple kinematic structure.
The construction of the proposed gravity compensators was realized by combining planar four-bar mechanisms with a rotating mass
and attaching a linear spring to each mechanism, then permuting the springs. An optimization procedure was exploited to determine
the link dimensions and spring parameters of the gravity compensators. Performance analysis and experiments with the gravity
compensators demonstrated their design feasibilities and good performances in which a torque reduction rate of 87.8% was practically
obtained. Lastly, an application of the proposed gravity compensators to serial robots was described. The results illustrated that the
actuator torque of a serial robot over a prescribed workspace could be reduced by 98.2%. A prototype of a serial robot was also made to
demonstrate the applicability of the gravity compensation design.

25
V.L. Nguyen
Table 10
A summary of the actuator torques and torque reduction rates of the KUKA robot with different payloads.
Joints Outputs Payloads mc3
0 kg (–100%) 15 kg (–50%) 30 kg* (+0%) 45 kg (+50%) 60 kg (+100%) 75 kg (+150%) 90 kg (+200%)
26

J. O1 (shoulder) Max. compensated torque (N-m) 951 807.3 663.7 702.9 940.6 1198.6 1456.6
Torque reduction rate 84.7% 87% 89.3% 88.7% 84.8% 80.7% 76.5%
J. O2 (elbow) Max. compensated torque (N-m) 445.2 225.1 81.7 287.7 521.6 755.6 989.7
Torque reduction rate 77.7% 88.7% 95.9% 85.6% 73.9% 62.2% 50.5%

*The original payload (30 kg) was adopted for the gravity compensation design for the robot. The maximum uncompensated torques at joints O1 and O2 are 6204 and 2000.3 N-m, respectively.

Mechanism and Machine Theory 172 (2022) 104770


V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

Declaration of Competing Interest

The author declares that he has no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgments

This work was supported by the Ministry of Science and Technology (MOST), Taiwan (Grant No. MOST 110-2222-E-167-004).

Supplementary materials

Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.mechmachtheory.2022.
104770.

Appendix A. Kinematic model of the four-bar (RRRR) mechanism

As shown in Fig. 3, a vector loop equation for the four-bar (RRRR) mechanism is written as:
vA + vAB + vBC − vC = 0 (A1)

where
vAB = vB − vA , vBC = vC − vB (A2)
The vector loop equation (Eq. (A1)) can be transformed into an algebraic equation, i.e.:

l1 eiπ + l2 eiα + l3 eiβ − l4 ei(θ+σ) = 0 (A3)

or equivalently
1 ( i(θ+σ) )
eiβ = l4 e − l2 eiα − l1 eiπ (A4)
l3
Eq. (A4) can be separated into a real part and an imaginary part, i.e.:
1
cosβ = [l4 cos(θ + σ ) − l2 cosα + l1 ] (A5)
l3

1
sinβ = [l4 sin(θ + σ) − l2 sinα] (A6)
l3
By relating Eqs. (A5) and (A6) with the equality (sin2β + cos2β = 1) to cancel the angle β, it can yield:
κA cosα + κB sinα + κC = 0 (A7)

where
2l2 2l2 l4 1[ ]
κA = − [l4 cos(θ + σ ) + l1 ], κB = − 2 sin(θ + σ), κC = 2 l24 + l22 + l21 + 2l1 l4 cos(θ + σ ) − 1 (A8)
l23 l3 l3
By using the trigonometric identities: sinα = 2ρ/(1+ρ2) and cosα = (1–ρ2)/(1+ρ2) with ρ = tan(α/2), Eq. (A7) can be rewritten as a
quadratic form, i.e.:

(κC − κA )ρ2 + 2κB ρ + κA + κC = 0 (A9)


Eq. (A9) can be solved to give the quantity ρ. Then, the joint angle α is obtained as:
( √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅)
− 2κB ± 4κ2B − 4(κC − κA )(κA + κC )
α = 2tan− 1 (A10)
2(κC − κA )

Eq. (A10) indicates that the polar angle of link 2 (α) can be determined once that of link 4 (θ) is given.

Appendix B. Seven algebraic equations arranged into a linear matrix equation (Eq. (36)) for type F

− F1x + F2x = 0 (B1)

− F1y − m2 g + F2y = 0 (B2)

27
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

F2x + F4x − Fsx = 0 (B3)

F2y + F4y − Fsy + m3 g = 0 (B4)

(xB − xA )F2y − (yB − yA )F2x − (xP2 − xA )m2 g + Ta2 = 0 (B5)

(xB − xA )F1y + (yA − yB )F1x − (xP2 − xB )m2 g − Tr3 = 0 (B6)

(xD − xB )Fsy − (yD − yB )Fsx − (xC − xB )F4y + (yC − yB )F4x − (xP3 − xB )m3 g + Tr3 = 0 (B7)

Appendix C. Particle swarm optimization (PSO) algorithm

Particle swarm optimization (PSO) is a metaheuristic algorithm as a stylized representation of the flocking behavior of birds [52,
53]. In PSO, the movement of each particle (candidate solution) in the search space is guided by its local best position and the global
best position of the entire swarm (population N). More specifically, the PSO algorithm is described through six steps, as detailed below:

1 Step 1. Initiate an array of random particles with consideration of the initial constraints.
2 Step 2. Evaluate the objective function.
3 Step 3. Find the local and global best positions based on the previous best positions.
4 Step 4. Update the velocity and position of the particles in the m-dimensional space by using Eqs. (C1) and (C2) below. The velocity
and position of a new particle [Vi(t + 1), Xi(t + 1)] are calculated from the velocity and position of the preceding particle [Vi(t),
Xi(t)] as:
Vi (t + 1) = Vi (t) + ε1 c1 [Pi (t) − Xi (t)] + ε2 c2 [Gi (t) − Xi (t)] (C1)

Xi (t + 1) = Xi (t) + Vi (t + 1) (C2)

where i and t represent the particle and discrete-time indices, respectively; ε1 and ε2 random values in the range (0, 1); c1 and c2
positive constants used for adjusting the speed of particle flying; Pi(t) and Gi(t) the local and global best positions, respectively.

1 Step 5. Update the local and global best positions.


2 Step 6. Return to Step 2 and repeat if a termination criterion is not satisfied.

Appendix D. Differential evolution (DE) algorithm

Differential evolution (DE) is a parallel direct search optimization method that performs through four stages: initialization, mu­
tation, crossover, and selection [54,55]. More specifically, the four stages of the DE algorithm are described as follows:

1 Stage 1. Initialization of a population of n d-dimensional random value vectors Xi,G (i = 1, 2,…, n) for each generation G (G = 0, 1,
…, Gmax). Each initiated vector is a genome that provides a candidate solution for the optimization problem. The initial population
(G = 0) is randomly generated to cover the entire parameter space.
2 Stage 2. Mutation with difference vectors to generate new parameter vectors. This operation is achieved by adding the weighted
difference between two population vectors to a third vector. For each target vector Xi,G (i = 1, 2,..., n), a generated mutant vector is
determined as:
( )
Vi,G+1 = Xr1 ,G + FC Xr2 ,G − Xr3 ,G (D1)

where r1, r2, and r3 represent the random indexes (r1, r2, r3 ∈ {1, 2,..., n}), and they should be chosen to be different from the running
index i; FC a scaling factor (FC ∈ [0, 2]) to control the amplification of the differential variation (Xr2,G – Xr3,G).

3 Stage 3. Crossover for parameter mixing to increase the diversity of the perturbed parameter vectors. The mutant vector’s pa­
rameters are mixed with the parameters of the target vector Xi,G to derive a trial vector Ui,G+1 = (U1i,G+1, U2i,G+1,..., Udi,G+1). Each
component Uji,G+1 (j = 1, 2,…, d) of the trial vector Ui,G+1 is expressed as:
{
Vji,G+1 if randu(j) ≤ CR or j = rnbr(i)
Uji,G+1 = (D2)
Xji,G+1 otherwise

28
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

where randu(j) stands for a uniformly distributed random number for each component j of the ith parameter vector (randu(j) ∈ [0, 1]);
CR the user-defined crossover factor (CR ∈ [0, 1]); rnbr(i) a randomly selected index (rnbr(i) ∈ {1, 2,…, d}) used to ensure that the trial
vector Ui,G+1 receives at least one parameter from the mutant vector Vi,G+1.

4 Stage 4. Selection of the target vector for the next generation. The trial vector Ui,G+1 is compared to the target vector Xi,G to define
the target vector Xi,G+1 for the next generation (G+1). The selection operation for the target vector Xi,G+1 is expressed as:
{
Ui,G+1 if Ui,G+1 ≤ Xi,G
Xi,G+1 = (D3)
Xi,G otherwise

References

[1] C. Gosselin, Gravity compensation, static balancing and dynamic balancing of parallel mechanisms. Smart Devices and Machines for Advanced Manufacturing,
Springer-Verlag London Limited, 2008, pp. 27–48.
[2] V. Arakelian, S. Briot, Balancing of linkages and robot manipulators: advanced methods with illustrative examples. Mechanisms and Machine Science, Springer
International Publishing, Switzerland, 2015.
[3] M. Carricato, C. Gosselin, A statically balanced Gough/Stewart-type platform: conception, design, and simulation, ASME J. Mech. Robot. 1 (3) (2009), 031005.
[4] V. Arakelian, Gravity compensation in robotics, Adv. Robot. 30 (2) (2016) 79–96.
[5] D. Ludovico, P. Guardiani, F. Lasagni, J. Lee, F. Cannella, D.G. Caldwell, Design of non-circular pulleys for torque generation: a convex optimisation approach,
IEEE Robot. Autom. Lett. 6 (2) (2021) 958–965.
[6] J.L. Herder, “Energy-free systems: theory, conception and design of statically balanced spring mechanisms,” Ph.D. Thesis, department of design, engineering and
production, Delft University of Technology, Delft, The Netherlands, (2001).
[7] C.M. Gosselin, J. Wang, On the design of gravity-compensated six-degree-of-freedom parallel mechanisms, in: Proceedings of the IEEE International Conference
on Robotics and Automation (ICRA), Leuven, Belgium, May 20, 1998, pp. 2287–2294.
[8] V. Van der Wijk, Design and analysis of closed-chain principal vector linkages for dynamic balance with a new method for mass equivalent modeling, Mech.
Mach. Theory 107 (2017) 283–304.
[9] J.M. Audet, C. Gosselin, Rotational low-impedance physical human–robot interaction using underactuated redundancy, ASME J. Mech. Robot. 13 (1) (2021),
014503.
[10] J.M. Audet, C. Gosselin, Intuitive physical human-robot interaction using an underactuated redundant manipulator with complete spatial rotational capabilities,
ASME J. Mech. Robot. 14 (1) (2022), 011011.
[11] V. van der Wijk, The Grand 4R four-bar based inherently balanced linkage architecture for synthesis of shaking force balanced and gravity force balanced
mechanisms, Mech. Mach. Theory 150 (2020), 103815.
[12] T. Essomba, Design of a five-degrees of freedom statically balanced mechanism with multi-directional functionality, Robotics 10 (1) (2021) 11.
[13] V.L. Nguyen, C.Y. Lin, C.H. Kuo, Gravity compensation design of Delta parallel robots using gear-spring modules, Mech. Mach. Theory 154 (2020), 104046.
[14] C.H. Kuo, V.L. Nguyen, D. Robertson, L.T. Chou, J.L. Herder, Statically balancing a reconfigurable mechanism by using one passive energy element only: a case
study, ASME J. Mech. Robot. 13 (4) (2021), 040904.
[15] J. Wang, X. Kong, A geometric approach to the static balancing of mechanisms constructed using spherical kinematic chain units, Mech. Mach. Theory 140
(2019) 305–320.
[16] P.Y. Lin, W.B. Shieh, D.Z. Chen, Design of statically balanced planar articulated manipulators with spring suspension, IEEE Trans. Robot. 28 (1) (2012) 12–21.
[17] H. Jamshidifar, A. Khajepour, T. Sun, N. Schmitz, S. Jalali, R. Topor-Gosman, J. Dong, A novel mechanism for gravity-balancing of serial robots with high-
dexterity applications, IEEE Trans. Med. Robot. Bion. 3 (3) (2021) 750–761.
[18] D. Lee, T. Seo, Lightweight multi-DOF manipulator with wire-driven gravity compensation mechanism, IEEE/ASME Trans. Mechatron. 22 (3) (2017)
1308–1314.
[19] H.S. Kim, J.B. Song, Multi-DOF counterbalance mechanism for a service robot arm, IEEE/ASME Trans. Mechatron. 19 (6) (2014) 1756–1763.
[20] W.B. Lee, S.D. Lee, J.B. Song, Design of a 6-DOF collaborative robot arm with counterbalance mechanisms, in: Proceedings of the IEEE International Conference
on Robotics and Automation (ICRA), Singapore, 2017, pp. 3696–3701. May 29 – June 3.
[21] T. Nakayama, Y. Araki, H. Fujimoto, A new gravity compensation mechanism for lower limb rehabilitation, in: Proceedings of the IEEE International Conference
on Mechatronics and Automation, Changchun, China, 2009, pp. 943–948. August 9–12.
[22] G. Endo, H. Yamada, A. Yajima, M. Ogata, S. Hirose, A passive weight compensation mechanism with a non-circular pulley and a spring, in: Proceedings of the
IEEE International Conference on Robotics and Automation (ICRA), Anchorage, AK, USA, 2010, pp. 3843–3848. May 3–7.
[23] H.S. Kim, J.K. Min, J.B. Song, Multiple-degree-of-freedom counterbalance robot arm based on slider-crank mechanism and bevel gear units, IEEE Trans. Robot.
32 (1) (2016) 230–235.
[24] M. Tschiersky, E.E. Hekman, D.M. Brouwer, J.L. Herder, Gravity balancing flexure springs for an assistive elbow orthosis, IEEE Trans. Med. Robot. Bion. 1 (3)
(2019) 177–188.
[25] T.Y. Tseng, Y.J. Lin, W.C. Hsu, L.F. Lin, C.H. Kuo, A novel reconfigurable gravity balancer for lower-limb rehabilitation with switchable hip/knee-only exercise,
ASME J. Mech. Robot. 9 (4) (2017), 041002.
[26] L. Zhou, W. Chen, W. Chen, S. Bai, J. Zhang, J. Wang, Design of a passive lower limb exoskeleton for walking assistance with gravity compensation, Mech. Mach.
Theory 150 (2020), 103840.
[27] A. Alamdari, R. Haghighi, V. Krovi, Gravity-balancing of elastic articulated-cable leg-orthosis emulator, Mech. Mach. Theory 131 (2019) 351–370.
[28] C.H. Kuo, S.J. Lai, Design of a novel statically balanced mechanism for laparoscope holders with decoupled positioning and orientating manipulation, ASME J.
Mech. Robot. 8 (1) (2016), 015001.
[29] C.K. Kim, D.G. Chung, M. Hwang, B. Cheon, H. Kim, J. Kim, D.S. Kwon, Three-degrees-of-freedom passive gravity compensation mechanism applicable to
robotic arm with remote center of motion for minimally invasive surgery, IEEE Robot. Autom. Lett. 4 (4) (2019) 3473–3480.
[30] T. Rahman, R. Ramanathan, R. Seliktar, W. Harwin, A simple technique to passively gravity-balance articulated mechanisms, ASME J. Mech. Des. 117 (4) (1995)
655–658.
[31] N. Ulrich, V. Kumar, Passive mechanical gravity compensation for robot manipulators, in: Proceedings of the IEEE International Conference on Robotics and
Automation (ICRA), Sacramento, CA, USA, 1991, pp. 1536–1541. April 9–11.

29
V.L. Nguyen Mechanism and Machine Theory 172 (2022) 104770

[32] D. Fedorov, L. Birglen, Differential noncircular pulleys for cable robots and static balancing, ASME J. Mech. Robot. 10 (6) (2018), 061001.
[33] V. Arakelian, Y. Zhang, An improved design of gravity compensators based on the inverted slider-crank mechanism, ASME J. Mech. Robot. 11 (3) (2019),
034501.
[34] V.L. Nguyen, C.Y. Lin, C.H. Kuo, Gravity compensation design of planar articulated robotic arms using the gear-spring modules, ASME J. Mech. Robot. 12 (3)
(2020), 031014.
[35] K. Koser, A cam mechanism for gravity-balancing, Mech. Res. Commun. 36 (4) (2009) 523–530.
[36] B.G. Bijlsma, G. Radaelli, J.L. Herder, Design of a compact gravity equilibrator with an unlimited range of motion, ASME J. Mech. Robot. 9 (6) (2017), 061003.
[37] G. Lee, D. Lee, Y. Oh, One-piece gravity compensation mechanism using Cam mechanism and compression spring, Int. J. Precis. Eng. Manuf. Green Technol. 5
(3) (2018) 415–420.
[38] N. Takesue, T. Ikematsu, H. Murayama, H. Fujimoto, Design and prototype of variable gravity compensation mechanism (VGCM), J. Robot. Mechatron. 23 (2)
(2011) 249–257.
[39] Y.C. Hung, C.H. Kuo, A novel one-DoF gravity balancer based on cardan gear mechanism, mechanisms and machine science. New Trends in Mechanism and
Machine Science, Springer International Publishing Switzerland, 2017, pp. 261–268.
[40] K.H. Ahn, W.B. Lee, J.B. Song, Reduction in gravitational torques of an industrial robot equipped with 2 DOF passive counterbalance mechanisms, in:
Proceedings of the IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS), Daejeon, South Korea, 2016, pp. 4344–4349. October 9–14.
[41] W.B. Shieh, B.S. Chou, A novel spring balancing device on the basis of a scotch yoke mechanism, in: Proceedings of the 14th IFToMM World Congress, Taipei,
Taiwan, 2015, pp. 206–212. October 25–30.
[42] J. Zhang, A.D. Shaw, M. Amoozgar, M.I. Friswell, B.K. Woods, Bidirectional spiral pulley negative stiffness mechanism for passive energy balancing, ASME J.
Mech. Robot. 11 (5) (2019), 054502.
[43] J.A. Franco, J.A. Gallego, J.L. Herder, Static balancing of four-bar compliant mechanisms with torsion springs by exerting negative stiffness using linear spring
at the instant center of rotation, ASME J. Mech. Robot. 13 (3) (2021), 031010.
[44] Q. Xie, S. Liu, H. Jiang, Design of a passive constant-force mechanism based on a five-bar mechanism, Mech. Mach. Theory 143 (2020), 103662.
[45] M. Li, W. Cheng, R. Xie, Design and experimental validation of two cam-based force regulation mechanisms, ASME J. Mech. Robot. 12 (3) (2020), 031003.
[46] S. Sanchez-Salinas, C. Nunez-Torres, J. Lopez-Martinez, D. Garcia-Vallejo, J.M. Muyor, Design and analysis of a constant-force bench press, Mech. Mach. Theory
142 (2019), 103612.
[47] F. Ma, G. Chen, H. Wang, Large-stroke constant-force mechanisms utilizing second buckling mode of flexible beams: evaluation metrics and design approach,
ASME J. Mech. Des. 142 (10) (2020), 103303.
[48] Y. Wei, Q. Xu, Design of a new passive end-effector based on constant-force mechanism for robotic polishing, Robot. Comput. Integr. Manuf. 74 (2022), 102278.
[49] B. Ding, J. Zhao, Y. Li, Design of a spatial constant-force end-effector for polishing/deburring operations, Int. J. Adv. Manuf. Technol. 116 (2021) 3507–3515.
[50] P. Wang, Q. Xu, Design and modeling of constant-force mechanisms: a survey, Mech. Mach. Theory 119 (2018) 1–21.
[51] R.G. Budynas, J.K. Nisbett, Shigley’s Mechanical Engineering Design, 9th ed., The McGraw-Hill Companies, Inc., New York, NY, USA, 2011.
[52] R. Eberhart, J. Kennedy, A new optimizer using particle swarm theory, in: Proceedings of the Sixth International Symposium on Micro Machine and Human
Science, Nagoya, Japan, 1995, pp. 39–43. October 4–6.
[53] Z. Meng, G. Li, X. Wang, S.M. Sait, A.R. Yıldız, A comparative study of metaheuristic algorithms for reliability-based design optimization problems, Arch.
Comput. Methods Eng. 28 (2021) 1853–1869.
[54] R. Storn, K. Price, Differential evolution–a simple and efficient heuristic for global optimization over continuous spaces, J. Glob. Optim. 11 (1997) 341–359.
[55] S. Das, P.N. Suganthan, Differential evolution: a survey of the state-of-the-art, IEEE Trans. Evol. Comput. 15 (1) (2011) 4–31.
[56] K.R. GmbH, The KR QUANTEC Series, KR 210 R3100 Ultra,” KUKA Industrial Robotics_High Payloads, KUKA Deutschland GmbH, Zugspitzstrasse 140 (2019)
86165.
[57] V.L. Nguyen, C.H. Kuo, P.T. Lin, Gravity balancing reliability and sensitivity analysis of robotic manipulators with uncertainties, in: Proceedings of the ASME
International Design Engineering Technical Conferences and Computers and Information in Engineering Conference (IDETC/CIE ), 2021. Online, Virtual, August
17–20DETC2021-66762.
[58] W.D. Van Dorsser, R. Barents, B.M. Wisse, J.L. Herder, Gravity-balanced arm support with energy-free adjustment, ASME J. Med. Devices 1 (2) (2007) 151–158.
[59] Z.W. Yang, C.C. Lan, An adjustable gravity-balancing mechanism using planar extension and compression springs, Mech. Mach. Theory 92 (2015) 314–329.
[60] J. Kim, J. Moon, J. Kim, G. Lee, Compact variable gravity compensation mechanism with a geometrically optimized lever for maximizing variable ratio of torque
generation, IEEE/ASME Trans. Mechatron. 25 (4) (2020) 2019–2026.

30

You might also like