You are on page 1of 17

Mechanism and Machine Theory 174 (2022) 104889

Contents lists available at ScienceDirect

Mechanism and Machine Theory


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

A novel design of planar high-compliance joint in variable stiffness


module with multiple uniform stress leaf branches on
rigid-flexible integral linkage
Fanghua Mei a, Shusheng Bi a, Linkun Chen a, Hanjun Gao a, b, *
a
School of Mechanical Engineering and Automation, Beihang University, Beijing 100191 China
b
Jingdezhen Branch of Jiangxi Research Institute, Beihang University, Jiangxi 333000 China

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

Keywords: Benefit from the elastic element in the drivetrain, the variable stiffness actuators offer advantages
Planar torsion spring over traditional rigid actuators in security, robustness, energy consumption. However, it is also
High-compliance very bulky due to the integration of the elastic element, especially with large deflection and high
Rigid-flexible integral linkage
load. This article introduces a novel planar high-compliance joint, in which the torsion spring is
Cantilever leaf feature
Uniform stress leaf branch
not an independent part but a synergy result of two adjacent parts developed from the stiffness
Analytical expression adjustment linkage mechanism. The flexible element is a cantilever leaf feature belongs to the
rigid-flexible integral linkage. The deflection and load ability of the torsion spring can be
enhanced by arranging multiple uniform stress leaf branches on the integral linkage without
increasing the size. Explicit analytical formulas are derived in a form convenient for leaf pa­
rameters design and verified by the finite element simulation and prototype experiments. Finally,
the error factors are quantitatively analyzed, which can provide information for the leaf
correction design. This kind of planar high-compliance joint is expected to be an ideal choice for
the variable stiffness actuators that require stringent axial compactness with large deflection and
high load.

1. Introduction

Based on the traditional rigid actuator, the flexible actuator is coupled to the load via a flexible element [1,2], which turns the force
control problem into a position control one. It brings great advantages in force control accuracy, stability [3,4] and friendly
human-machine interaction [5–8]. And its advantages have been greatly expanded by the adjustable stiffness module [9]. Unfortu­
nately, there is no large-scale commercial application, and the insufficient compactness of the flexible element is one of the key factors,
especially with large deflection and high load. It is hard to tackle the contradiction between the deflection and load in a compact way.
According to its implementation principle, the flexible element in the literatures can be divided into three categories: the preload
spring based on conventional spring, the leaf spring based on beam theory with small deflection, the flexure hinge based on nonlinear
model undergoing large deformation. Tsagarakis et al. [10] achieved a rotary spring based on a novel arrangement of six linear springs
with three spokes structure, and Vo et al. [11] changed the springs position with three sets of synchronous stiffness adjustment linkage
mechanism. Wolf et al. [12], Yigit et al. [13], Guo [14] and Ayoubi et al. [15] used the cam to decompose the linear spring

* Corresponding author at: Hai-Dian District Xue Yuan Road 37 #, Beijing, China.
E-mail address: hjgao@buaa.edu.cn (H. Gao).

https://doi.org/10.1016/j.mechmachtheory.2022.104889
Received 11 January 2022; Received in revised form 2 April 2022; Accepted 18 April 2022
Available online 27 April 2022
0094-114X/© 2022 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

deformation, and converted the axial linear spring into torsion spring in a more compact way and its modular design is intended for
application in multi-degree-of-freedom robotic systems [16,17]. Jafari et al. [18] and Jin et al. [19] achieved the elastic element by
two opposite-handed helical torsion springs and two cam-bearing followers. For the linear spring or helical torsion spring structure, it
is convenient to design and manufacture, while requires complex mechanism to convert their operation mode to rotational motion [20,
21] and needs to preload the spring for compliance in both directions [22,23].
The leaf spring has the characteristics of easy integration and no internal stress. Furnémont et al. [24] and Shao et al. [25] stacked
spiral spring along the joint axis direction, allowing a more compact spring in comparison to the previous preload spring. Negrello et al.
[26] improved the deformation ability by maximizing the effective length of the beam through a series of radial and tangential
segments and reducing the beam end constraints in the spoke spring. Xu et al. [27] designed a compact S-shaped springs with a variable
cross-section beam. Based on the beam theory, analytical equations are available in a form useful for spring design in small deflection
(max deflection less than 1/10 the span). For the case with gradually variable thickness leaf [28] and long effective length over its
height or width (large than 10), the Euler beam theory can provide an acceptable accuracy, which is also verified in many leaf-type
flexure hinges with constant or variable thickness [29–33]. However, the deflection and load ability are limited and needs to be
enhanced by arranging multiple leaves in radial direction [34–36] (or in axial direction [37]), which deteriorates the compactness.
The flexure-hinge can realize a large angular deflection stroke with special-shaped flexible beam [38] and directly embed the
flexible element on the arbitrary position of the kinematic chain without the bearing [39]. For such flexible hinges, its effective elastic
segments are intended to undergo geometric nonlinearities. From the perspective of reducing modeling error, the existing integration
methods based on Euler theory show some limitations [40]. The deflection modeling should be studied by some complicate but more
accurate methods such as chained beam-constraint-model (CBCM) [41], pseudo-rigid model [42] and finite element method [43,44].
During the modeling process of CBCM, the beam is discretized equally into N elements and CBCM equations can be obtained and
numerically solved based on Bernoulli–Euler theory [45]. For the building of the pseudo-rigid model, flexure hinges are replaced by
revolute joints and lumped torsional springs [46]. And then the well-known kinematic and dynamic methodologies for both synthesis
and analysis can be applied [47]. In order to increase the accuracy, pseudo-rigid mechanisms may include several links and joints,
especially for the building of mechanisms with several flexure hinges [48,49]. Palli et al. [50] achieved a flexible hinge with
approximately quadratic mechanical characteristics by compliant four-bar linkage mechanism. Bilancia et al. [51] designed a high
compliant element with quadratic torque-deflection relationship by slender spline beams. Aiple et al. [52] changed the cross-section
shape of the winding spring for the desired deflection at maximum load by finite element simulation. The shape and dimensions of
flexure spring needed to be defined through an iterative optimization process without analytical expression [53].
For the flexible element design, there is no winner, but rather principle-dependent optimal solution. In this paper, a new planar
high-compliance joint is presented based on its stiffness adjustment linkage mechanism with the following advantages:

(1) Rigid-flexible integral component design for high compactness. The stiffness adjustment module is realized based on the guide
rod mechanism. The paper makes a full utilization of the guide rod mechanism. The torsion spring is not an additional inde­
pendent part but a synergistic result of two adjacent parts developed from the linkage mechanism, in which the flexible element
is a cantilever leaf feature belongs to the developed rigid-flexible integral linkage. So no more structural space is consumed as
the torsion spring is performed by adding some features on the original linkage components.
(2) Multiple uniform stress leaf branches on the integral linkage for higher deflection. The deflection enhancement is performed in a
compact way by arranging multiple uniform stress leaf branches on the linkage for a smaller thickness, without the complicated
leaves synchronous mechanism or leaves stack configuration. The neutral axes of all branches are concurrent at the compliant
joint for force symmetry.
(3) Explicit analytical expression convenient for parameters design with large deflection angle. The deflection angle on the joint is
enlarged multiple times in the deformation mode conversion process, which can ensure the actual deflection of the leaf conform
to the small deformation hypothesis of the beam theory. There are explicit analytical expressions among the leaf parameters and
torsion spring performance in large deflection and the performance can be easily scaled up to meet the requirements by
changing the design parameters.

The following content mainly includes four parts. The Section 2 mainly introduces the design principle and requirements of the
planar torsion spring based on guide rod mechanism using cantilever leaf feature. TheSection 3 mainly introduces the design method of
multiple uniform stress leaf branches for deflection enhancement and establishes the analytical expressions among the leaf parameters
and its performance indexes. The Section 4 makes a comparison analysis between the finite element simulation result and the

Fig. 1. . Basic components of stiffness adjustment module.

2
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

prototype measurement result, quantitatively analyses the error factors of torsion spring design results.

2. Design of compliance joint based on guide rod mechanism

2.1. Principle of planar bidirectional torsion spring design

The stiffness adjustment module is based on a classic guide rod mechanism, which includes link OA, link AC, link OB, the hinged
joints A and B, pivot O, sliding joint C (Fig. 1). The position of joint B is variable both in axial and radial direction [54], but in this paper
the link OB is set as a fixed body for simple analysis. The kinematics relationship between the rotary angle of link OA and its trans­
mission angle (θ and α in Fig. 1) is decided by Eq. (1), i.e., transmission ratio dα/dθ is variable by adjusting the length of link OB.
θ = α + arcsin(OAsinα/OB) (1)
The elastic element is arranged on the joint A as a bidirectional torsion spring and the original guide rod mechanism is then
converted to a variable stiffness module. Now the angle θ and α are respectively called as the deflection angle of the stiffness
adjustment module and torsion spring, and the hinged joint A can be called as compliant joint. Based on principles of virtual work, the
statics equations between the torque on link OA and torque induced by torsion spring deflection (τOA and τs in Fig. 1) are obtained as:


⎨ τOA =∫ τs dα/dθ
α
(2)
⎩ τs =
⎪ ks dα
0

where ks is the stiffness of the torsion spring. The stiffness of the stiffness adjustment module (KO in Fig. 1) can be defined as:
( )2
dτOA dα d2 α
KO = = ks + τs 2 (3)
dθ dθ d θ

KO is determined by torsion spring stiffness (ks) and transmission ratio (dα/dθ). The compactness of the torsion spring directly decides
the size of the stiffness adjustment module. To achieve a planar torsion spring on joint A without increasing the size, the rigid link OA
(red colored link in Fig. 2a) is developed to a new rigid-flexible integral component (also marked red in Fig. 2b). The rigid element is
equal to the original link OA. The flexible element is a cantilever leaf feature belongs to the new component. When the rigid-flexible
integral component revolves around pivot O, the deflection of the leaf is converted to deflection angle on the linkage joint A (w and α in
Fig. 2b) through the roller D on the adjacent link AC. Under the synergy of the leaf and the roller D, the equivalent plane bidirectional
torsion spring is embedded on joint A without any new independent part (marked red in Fig. 2c).
The equivalent deflection angle on the joint A is defined as:
α = w/AD (4)

The ε represents the leaf end slope (Fig. 2b) and is related to w. The Fs(w) represents the reaction force on the leaf when the leaf end
deflection is w. The equivalent torque on the torsion spring is given by:
τs = Fs (w)cos(ε(w) + α(w))AD (5)

As all the variables in Eqs. (4) and (5) can be determined directly by deflection w, it is obviously easier to derive the torsional spring
stiffness using the w as intermediate variable. Thus:

Fig. 2. Design principle of planar bidirectional torsion spring.

3
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

∂τs ∂w
ks = (6)
∂w ∂α

2.2. Constraints and requirements of torsional spring design

In the specific structure design, a rigid U-shaped chute needs to be set at the end of the cantilever leaf for bidirectional deform­
ability, and bearings are installed on the roller D to reduce the elastic hysteresis caused by internal friction, as shown in the Fig. 3. llf is
the leaf effective length, lrd is the length of the rigid U-shaped chute, b is the leaf width, h(x) is the leaf thickness at the position x, hr is
the leaf thickness at the leaf root.
Considering the specific leaf structure in Fig. 3, the operation mode conversion between the leaf (reaction force Fs, slope ε,
deflection w) and the torsion spring is shown in Fig. 4a. Its equivalent torque τs, deflection angle α is shown in Fig. 4b. lrod is the length
of AD.
Performance index requirements. The maximum deflection angle and bearing capacity of the torsional spring directly determines
the performance of the stiffness adjustment module, which is the core of the leaf design. The performance indexes are that:α is not less
than 0.22 rad, and τs is not less than 24 Nm.
Leaf stress constraint. According to Eq. (3), the nonlinearity of the stiffness adjustment module is determined by ks and dα/dθ.
Among them, nonlinearity caused by the high-rate plastic deformation of spring material is difficult to calibrate and compensate
precisely. A simple way to avoid this nonlinearity is keeping the stress-strain in elastic behavior in the whole deflection process.
Proportional limit is the point on a stress-strain curve at which it begins to deviate from the straight-line relationship between the stress
and strain, and the corresponding bending stress is defined as σp. The maximum α,τs of the torsion spring is defined as αp and τsp
respectively. The stress constraint on the leaf is that: the maximum stress σ max on the leaf should not exceed its material proportional
limit σ p.
Small deflection constraint. The max deflection w should be less than 1/10 the leaf span (0.22lrod≤llf+lrd), so that the formulas can
be derived accurately based on the Euler beam theory.
Strength constraint. The reduction of AD is conducive to increase α with a smaller deflection on the leaf, but it will also increase the
reaction force Fs on the bearings at the same time. The maximum reaction force should not exceed the bearing’s rated load, which is
[Fbear].
Dimension constraint. The parameters of the leaf structure should not lead to an increasing size of the stiffness adjustment module
(llf+lrd+lrod = 117.5 mm, b ≤ 7 mm).The design constraint equations of torsional spring are as follows:


⎪ α ≥ 0.22rad
⎪ τ ≥ 24Nm


⎪ s

⎪ σ max ≤ σp

Fs ≤ [Fbear ] (7)


⎪ llf + lrd + lrod/ =( 117.5mm

⎪ )

⎪ 0.22lrod ≤ 1 10 llf + lrd


b ≤ 7mm

3. Deflection enhancement design of the torsional spring using the cantilever leaf

3.1. Torsional spring design with uniform stress leaf

The maximum stress σ (x)max at the position x is:


( )
6Fs llf + lrd − x
σ(x)max = (7a)
bh2 (x)

Fig. 3. Variable cross-section cantilever leaf with rigid U-shaped groove.

4
F. Mei et al.
5

Fig. 4. (a) Conversion of leaf force-deflection to torque-deflection on joint A (b) Equivalent torsional spring on joint A.

Mechanism and Machine Theory 174 (2022) 104889


F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

To enhance the deflection angle αp, the cross-section thickness h(x) away from the leaf root should be decreased according the uniform
stress equation for evenly strained, as Eq. (8).
( )
h2 (x) = p llf + lrd − x (8)

where p is the uniform stress leaf thickness coefficient to be determined, and is the key parameter calculated by:
6Fs 6τs
p= = (9)
bσmax bσ max lrod

Based on the classical Euler Bernoulli beam theory, the deflection and slope of the leaf end can be given by:
⎧ ( ) [( )0.5 ]

⎪ ε llf = 2m lrd + llf − (lrd )0.5




⎨ ( ) 4 ( )0.5 4 ( )1.5
w llf = m(lrd )1.5 + 2m lrd + llf
3
llf − m lrd + llf
3 (10)





⎩ m = 12Fs

p1.5 Eb

Considering the rigid U-shaped groove, the deflection at the point D is:
( ) ( ) 8Fs (( )1.5 )
w = w llf + ε llf lrd = 1.5 lrd + llf − (lrd )1.5 (11)
p Eb

Going back to deflection angle Eq. (4), we obtain


w 4σ max (( )1.5 )
α= = 0.5 lrd + llf − (lrd )1.5 (12)
lrod 3p Elrod

The deformation rate is defined as the leaf deflection over its span by:
( )/ ( ( ) )/
δ = w llf llf = αlrod − ε llf lrd llf

To calculate p, we need to know Fs, b, σmax. The force Fs cannot be directly calculated by Eq. (5) due to the unknown variables (ε and α).
There is a coupling relationship among τs, Fs, w, and the above design calculation needs to be completed through multiple iterations.
This process is not conducive to the leaf parameters design. A simpler way is to given 0.5 times design margin on τs to cover the
influences of the cosine term cos(ε+α) for a concise analytical expression. Then the force Fs can be calculated by the simplified Eq. (13).
Fs = τs /lrod (13)

The simplified stiffness ks can be expressed as:

τs p1.5 Eb(lrod )2
ks = = [( ] (14)
α 8 lrd + llf )1.5 − (lrd )1.5

The Eq. (14) represents the torsion spring stiffness when α is nearly zero. As α increases, the accurate τs and ks should be recalculated
according to Eqs. (5) and (6) with the above parameters design result. The Eqs. (9) and (12)–(14) provide explicit mathematical
expressions between the design parameters (p, b, E, lrod, llf, lrd, σ max) and the performance indexes (τs, α, ks).
The leaf material selected is 60Si2Mn, which has small elastic hysteresis and high bending stress. Its bending stress limit σ b is 1274
MPa and Young’s mode E is 206,000 MPa. According to the engineering experience of the spring manufacturer, the material behavior
is proportional when the maximum stress is less than 0.75σ b. Under the constraints of leaf maximum stress and the design margin
value, we set σmax=σp = 955 MPa, τsp=1.5τs = 32Nm.
The bearing on roller D selected is NSK-NB700 with rated bearing capacity of 515 N, width of 3 mm and diameter of 15 mm.
According to Eqs. (7a) and (13),there is
⎧ /
⎨ lrod ≥ τsp [Fbear ] = 62mm
l ≤ 117.5/3.2 = 36.7mm
⎩ rod
llf = 117.5 − lrod − lrd

The bearing capacity of single bearing is too low, resulting in too large lrod, which cannot meet the small deflection constraint. Double
bearings are set on roller D to improve the bearing capacity for a smaller lrod and bigger llf. Thus, there is lrod = 31 mm, lrd = 9.5 mm, llf
= 77 mm. The minimum leaf width b is same to the axial size of the stiffness adjustment module, which is 7 mm. According to the
design requirements in Eq. (7a), the design results of the uniform stress leaf are as Table 1.
Considering the torque attenuation caused by the item cos(ε(w)+α (w)), the torsional spring torque τsp is 30.5 Nm according to Eq.
(5). When it is far from the leaf root, the leaf thickness is gradually reduced according to uniform stress Eq. (8), each cross-section of the

6
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

leaf is fully strained. However, the deflection angle αp is still insufficient, which is 23% less than the performance index 0.22 rad. In
order to furtherly improve the deflection angle αp, the possible methods are:

(1) Rechoosing a new leaf material. According to Eq. (12), it can be observed that α ∝ E− 1. We can rechoose a new leaf material with
smaller Young’s mode E, for example beryllium bronze (C17200) with E of 85 MPa and σ b of 1035 Mpa. If the pollution caused
by beryllium copper can be ignored, it will be a right choice with an appropriate leaf thickness coefficient p. Another common
material is hard aluminum alloy (2Al2-T4) with E of 70 MPa and σ b of 275 Mpa. But this would lead to a reduce in bearing ability
considering the effect of the lower σp. The performance of torsion spring is limited by the material mechanical properties.
(2) Directly decreasing the leaf thickness coefficient p. According to Eq. (12), it can also be observed that α ∝p− 0.5. The maximum α
will increase to 0.22 rad if the p decreases by 40.2%. The shortcoming is that this method will result in a significant increase in
the axis direction size of the stiffness adjustment module, as the width b increases by 0.67 times to maintain the same load
ability.
(3) Arranging more leaf branches on the rigid-flexible integral linkage. If N leaf branches experience the load τs together, the new
thickness coefficient can be roughly decreased to p/N according to Eq. (9), and correspondingly the deflection angle αp can be
nearly increased by N0.5–1 times. This method can reduce the requirements on the materials mechanical properties and exhibits
noticeable applicability to the requirements of different load levels. In this situation, a smaller lrod could be allowed since τs/N is
distributed to the single leaf branch, but its value is limited by the structure interference.

3.2. Torsion spring design with multiple uniform stress leaf branches on the rigid-flexible integral linkage

Theoretically, when the leaf has two completely symmetrical branches, each branch can evenly bear the load τs, the thickness
coefficient p can be reduced nearly to half of the original, and the maximum deflection angle α of the torsion spring can increase to
0.24 rad. In fact, the length of the leaf will be reduced under the dimension constraint in this case. To ensure sufficient design margin,
two branches are arranged on both sides of the middle branch to share the bending moment τs, as shown in Fig. 5. The neutral axes of
all branches are concurrent at joint A for force symmetry of the architecture. For the middle branch, the reaction force is Fm, the
equivalent torque and stiffness on the joint A is τm and km, the leaf effective length is lm, the leaf end slop is εm, the leaf thickness is hm
(hrm at leaf root), the thickness coefficient is pm, the distance from the center of the joint A to the center of the roller D1 is lrodm. In a
likewise fashion, the parameters and variables of the side branch are defined as Fp, τp, kp, lp, εp, hp, hrp, pp, lrodp, respectively.
The middle branch and side branches bear the load together, the simplified load formula is:
Fm lrodm + 2Fp lrodp = τs (15)

According to the uniform stress Eq. (9), the maximum stress σ max can be expressed by pm, pp, Fm and Fp as:
6Fm 6Fp
σmax = = (16)
bpm bpp

According to deflection angle Eq. (12), pm and pp meet the following deflection equation:
4σ max ( ) 4σmax (( )1.5 )
0.5
(lrd + lm )1.5 − (lrd )1.5 = 0.5
lrd + lp − (lrd )1.5 (17)
3pm Elrodm 3pp Elrodp

According to the Eqs. (16) and (17), pratio is defined as the ratio of the thickness coefficient between the middle branch and side branch,
and can be calculated as:
[( ) ]
Fm p m (lrd + lm )1.5 − (lrd )1.5 lrodp 2
pratio = = = (( ) ) (18)
Fp pp lrd + lp
1.5
− (lrd )1.5 lrodm

Combined with load Eq. (15), τm,τp can be expressed as:





τs lrodp
⎨ τp = Fp lrodp = pratio lrodm + 2lrodp

(19)

⎪ pratio τs lrodm

⎩ τm = Fm lrodm =
pratio lrodm + 2lrodp

Going back to Eq. (9) with τp and τm, the coefficient pm and pp can be calculated as:

Table 1
Torsional spring design performance with uniform stress leaf.
b lrod lrd llf τsp p hr ks (α=0) σp ε δ

7 mm 31 mm 9.5 mm 77 mm 30.5 Nm 0.825 mm 8.4mm 184Nm/rad 0.17 rad 0.14 rad 5.1%

7
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

Fig. 5. Rigid-flexible integral linkage with multiple leaf branches.


⎪ 6pratio τs lrodm

⎪ p = ( )
⎨ m σmax blrodm pratio lrodm + 2lrodp
(20)

⎪ 6τs lrodp

⎩ pp = ( )
σ max blrodp pratio lrodm + 2lrodp

The maximum deflection angle determined by Eq. (12) is:


4σmax ( )
α= 0.5
(lrd + lm )1.5 − (lrd )1.5 (21)
3(pm ) Elrodm

The leaf end slope εm and εp is:



⎪ 4σ [ ]

⎪ ε = max (lrd + lm )0.5 − (lrd )0.5
⎨ m p0.5 E
(22)
m


⎪ 4σmax [( )0.5 ]

⎩ εp = 0.5 lrd + lp − (lrd )0.5
pp E

The equivalent stiffness km and kp is:


⎧ (pm )1.5 Eb(lrodm )2

⎪ km = [ ]



⎨ 8 (lrd + lm )1.5 − (lrd )1.5
( )1.5 ( )2 (23)

⎪ pp Eb lrodp

⎪ [
⎩ kp = (
⎪ )1.5 ]
8 lrd + lp − (lrd )1.5

The stiffness of the torsional spring on joint A is:


ks = km + 2kp (24)

Similarly, the Eq. (24) represents the torsion spring stiffness when α is nearly zero and the accurate τs and ks must be recalculated
according to Eqs. (5) and (6) as α increases to αp. The results of parameters design and performance indexes are as Table 2 shows.
After introducing two side branches, the load carried by the single branch is reduced by nearly 2/3, allowing thickness coefficient of
the middle leaf branch to decrease by nearly 2/3. Benefit from this, deflection angle αp increases from 0.17 to 0.29 rad. The torsional
spring torque τsp is 28.14 Nm according to Eq. (5). In order to avoid the interference of the leaf branches, AD1 is 1.5 mm larger than AD.
The advantages of torsion spring using multiple uniform stress leaf branches are:

8
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

(1) The performance of the torsion spring can be scaled up by arranging more branches to match the design requirement. According
to Eqs. (16), (21) and (23), there is α ∝p− 0.5, p∝τs, ks∝p1.5. As αp is 31.8% larger than performance requirement, the thickness
coefficient pm and pp can be increased by 0.74 times, and the corresponding torsional spring torque τp and stiffness ks can be
increased by 0.74 and 1.29 times, respectively. There is a large improvement potential in the performance of the torsional spring
only by adjust the thickness coefficient, which brings greater flexibility to the design.
(2) The design margin can reduce the requirements on material mechanical properties. According to Eq. (21), there is α ∝σ max. The
max stress σ max on the leaf is 75.8% of its engineering experience value when the deflection angle α reaches 0.22 rad. This means
that the actual value of σp is allowed to be 24.2% smaller than its engineering experience value when the other material per­
formance parameters are the theoretical value. τsp is 17.25% larger than performance requirement, and correspondingly the
Young’s modulus E is allowed to be 17.25% smaller than its theoretical value.

The cross-section thickness curves of middle and side leaf branch are shown in Fig. 6.
According to the above thickness curves, the rigid-flexible integral component is designed and manufactured. After heat treatment
in a vacuum furnace, its Rockwell hardness is 38HRC, ensuring its ability to recover from deformation. The integrated structure of the
torsional spring is shown in the Fig. 7, which includes rigid-flexible integral component, link AC, bearings at joint A, bearings on
cylinders D1–D3, roller B, slider C and guide. The rigid-flexible integral component contains flexible element feature (side leaf branch
and middle leaf branch acting as a part of the torsion spring) and rigid element feature (acting as the original link OA). The link AC
includes four cylinders, in which the bearings fixed on the cylinder A forms linkage joint A between link AC and the rigid-flexible
integral component, and the bearings fixed on cylinder D1–D3 forms the rollers D1–D3 driving the leaf branches deform. The guide
is fixed on link AC, which allows the slider C moving along the link AC for a variable length of link OB. When the length of OB remains
unchanged, the rotation of the slider around the roller B can ensure the freedom required by the torsion spring’s deformation.

4. Experiment and result analysis

4.1. Validation method

Both the maximum deflection and the maximum load capacity test of the torsion spring may cause irreversible damage to the leaf.
Therefore, the results of torsion spring design parameters are verified by torque-deflection relationship measurement without reaching
its extreme performance. As the maximum value of ε+α on the middle branch and side branch are 0.52 rad and 0.48 rad respectively in
Table 2, the item cos(ε+α) in Eq. (5) can be re expressed as:

(ε + α)2 (ε + α)4
cos(ε + α) = 1 − + (25)
2 4!

Before reaching the point σp on the stress-strain curve, the nonlinearity is very weak and relatively simple, so the material behavior can
be expressed as:
∂F
= a0 + a1 w2 + a2 w4 (26)
∂w

The coefficient ai(i = 0,1,2) is decided by the leaf parameters. In additional, there is w = bα= lrodα, ε=cα, and the constant c is
determined by Eqs. (10) and (11). As the angle α increases, the τs in Eq. (5) can be expressed with different simplified equations in
different situations.
Situation 1: if α→0, then
{
τs = B1 α
B1 = a0 b2

Table 2
Torsional spring design performance with multiple leaf branches.
middle branch side branch single branch (in Section 3.1)

Thickness coefficient pm=0.281mm pp=0.354mm p = 0.83mm


Position of roller D AD1=32.5mm AD2=AD3=28mm AD=31mm
Length of the leaf branch lm=75.5mm lp=68.5mm llf =77mm
Stiffness (Nm/rad) km=36.43(α=0) kp= 35.28(α=0) ks=184(α=0)
Leaf load (N) Fm=334 Fp =376 Fs =1032
Leaf root thickness (mm) hrm=4.89 hrp=5.24 hr=8.4
Leaf end slope εm=0.23rad εp=0.19rad ε(llf)=0.135rad
Leaf deformation rate δ 9.6% 9.2% 5.1%
Deflection angle αp (rad) 0.29 0.29 0.17
Torsional spring torque τsp (Nm) 28.14 30.5
Leaf group stiffness ks (Nm/rad) 107(α=0) 184(α=0)

9
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

Fig. 6. Thickness curves of uniform stress leaf branches.

Fig. 7. Integrated prototype of the torsional spring.

Situation 2: if α→0.29 rad and ai (i = 1,2) is zero, then




⎪ τ s = B1 α + B2 α 3 + B3 α 5





⎪ B1 = a0 b2


a0 b2 (1 + c)2

⎪ B2 = −

⎪ 2



⎪ a b2
(1 + c)4

⎩ B3 = 0
4!

Situation 3: if α→0.22 rad and ai (i = 0,1,2) is non-zero constant, then




⎪ τ s = B1 α + B2 α 3 + B3 α 5





⎪ B1 = a0 b2


a0 b2 (1 + c)2 a1 b4

⎪ B2 = − +

⎪ 2 3



⎪ a b2
(1 + c) 4
a b4
(1 + c)2 a2 b5

⎩ B3 = 0 1
− +
4! 6 5

As the torque deflection relationship is an odd function, so the coefficient of the even term in the above torque expressions is zero.
Situation 1 is the case corresponding to Eq. (24); situation 2 can be performed with finite element software by setting the material
behavior as elastic; situation 3 can be obtained by the physical measurement of the prototype.

4.2. Stress observation and torque-deflection relationship experiment by finite element method

The finite element method (FEM) is an effective method to calculate and observe the stress distribution, strain, deformation on the

10
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

leaf and the total reaction torque around the joint A in detail. The leaf deformation simulation is performed by Abaqus software.
Material nonlinearity, geometric nonlinearity and boundary nonlinearity are the three sources of nonlinearity in structural me­
chanics simulations. The geometric nonlinearity is not obvious in this paper, as the elastic segment is a slender cantilever leaf with end
load and it requires that the leaf max deflection should be less than 1/10 its span to avoid the large deflection problem. Also, the
material nonlinearity is weak, for that the maximum leaf stress is designed to be smaller than its material proportional limit. But there
is boundary nonlinearity, as the contact area changes in the deformation process.
In the protype, the bearings are intentionally placed between the roller Di(i = 1,2,3) and the U-groove surfaces of the leaf branches.
The bearings will slide slightly along the surfaces when the leaf branches are deformed in the actual mechanism. So, the contact
relationship between the roller Di and the U-groove surfaces must be set as hard and frictionless surface-to surface contact with finite
sliding.
Deflection constraint ensures the applicability of Euler beam theory. Moreover, the influence of large deformation is not obvious, at
least for the case of the leaf dimension in this paper. In order to verify this conclusion, the mechanics simulations of the above middle
leaf branch are carried out with two different settings, one with the geometric nonlinearity choices on (Nlgeom on) and another off.
The vertical force is adjusted to 380 N as the load point is moved to the end of the leaf, for a clear observation of the deformation rate at
the leaf end. And its direction remains unchanged, always vertical down, as shown in Figs. 8a and 9a. The deflection of the former
(Nlgeom on) is only 1.6% smaller than that of the later (Nlgeom off) when deformation rate d reaches nearly 10%, as shown in Figs. 8a
and 9a. Due to the change of load point, the uniform stress no longer exists as expected (Figs. 8b and 9b).
The link AC is simplified as a rigid body, and all nodes on the rigid body are coupled to the center point of the hinged joint A. The
boundary conditions are that the outer surface of the rigid-flexible integral component is fixed, with U1=U2=U3=UR1=UR2=UR3=0;
the link AC is constrained, with U1=U2=U3=UR1=UR2=0, UR3=0.29 rad; the Nlgeom choice is turned on, allowing the finite sliding
among the rollers and U-shape leaf surfaces.
The general static solver cannot deal with the complex contact problem in this paper. Quasi-static analysis with ABAQUS/Explicit is
a true dynamic procedure originally developed to model high-speed impact events. But its accuracy is limited. For the low-speed
contact event in this paper, we treated it as a quasi-static one with the implicit dynamic solver for a higher accuracy. By setting the
mass of the mode to be nearly zero, inertial forces remain insignificant in the deformation process. Other parameters include the
Young’s modulus (206,000 MPa), Poisson’s ratio (0.3), material behavior (elastic behavior with yield stress of 1274 Mpa) and mesh
type (C3D20R). The simulation parameters setting is shown in Fig. 10.
The stress nephogram of leaf simulation result is shown in the Fig. 11. The maximum stress of the leaf is evenly distributed on the
side surface of the leaf branches, and its value is 910 MPa, which is very close to the proportional limit stress σ p of 60Si2Mn. The
reasons for the maximum stress discrepancy are as follows:

(1) The stress of U-shape groove is less than 300 MPa and there is a small deflection. In the theoretical mode, the U-shape groove is
treated as a completely rigid body.
(2) In the FEM mode, the roller has a small slip along the U-shape groove surface, which increases the actual length of the U-shape
groove. But in the theoretical mode, the effective length of the leaf is considered to remain unchanged in the whole deformation
process.

The above differences will reduce the leaf deflection and cause the maximum simulated stress on the leaf to be less than its
theoretical value. After the simulation, the reverse torque τs on the link AC around joint A and the corresponding rotation angle α are
output in Fig. 12. According to situation 2, a 5th-order fitting is carried out, and the boundary condition is that the coefficient of even
power series is zero, as Fig. 12 shows.
The polynomial coefficients fitted according to the Abaqus simulation result are marked with subscript a. According to the fifth-
order fitting curve equation of the simulation result, there is Ba1 = 103.9, Ba2 = − 127.7, Ba3 = 82.75. The first-order term
(103.9α) represents the torque-deflection relationship when α is nearly zero as situation 1. The third-order term (− 127.7α3) represents
the main effect of the cosine term cos(ε (w)+α (w)) on torque τs in Eq. (5), while the fifth-order term (82.75α5) has almost no effect on
the torque as its max value is 0.16 Nm. Therefore, the torque τsa and the stiffness ksa of torsion spring are fitted as:
{
τsa = − 127.7α3 + 103.9α
(27)
ksa = − 383.1α2 + 103.9

Fig. 8. Leaf simulation with Nlgeom on (a) strain nephogram;(b) Stress nephogram.

11
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

Fig. 9. Leaf simulation with Nlgeom off (a) strain nephogram;(b) Stress nephogram.

Fig. 10. Simulation parameters setting.

Fig. 11. Stress nephogram of leaf deformation.

The initial stiffness is ksa = 103.9 Nm/rad when the α is nearly zero. It is 2.9% smaller than the theoretical value (ks = 107 Nm/rad) in
Table 2. The maximum simulation torque is 27.05 Nm, which is 3.9% smaller than the theoretical value (τsp = 28.14 Nm) in Table 2.
The simulation error is very small as the simulation model is simple and close to the actual prototype. So, the simulation result can be
used as a substitute value for the theoretical value. The simulation verifies the correctness of the leaf design formulas from the aspects
of uniform stress nephogram, torque-deflection curve, maximum torque (α=0.29 rad), initial stiffness (α=0).

12
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

Fig. 12. Simulation and fitting results of torsion spring.

4.3. Torque-deflection indirectly measurement by prototype test

Due to structural reasons, it is difficult to measure the torque-deflection relationship of the torsion spring directly. But the
deflection angle and torque (θ and τOA in Fig. 1) of the stiffness adjustment module are easier to measure. Contrary to Fig. 1, the link OA
is fixed with the torque sensor on the frame, while the link OB is drived around the center pivot O by two motors (Fig. 13). The torque
τOA can be obtained by the torque sensor as a reverse torque and the deflection anlge θ can be measured by the magnetic mncoder on
the motor.
According to the Eqs. (1)–(3), if OA=OB, there would be:

⎨ θ = 2α
τ = τs /2 (28)
⎩ OA
KO = ks /4

Now, the torque-deflection curve measurement is converted from the torsion spring to the stiffness adjustment module, which can
obtain larger measurement angle, higher accuracy.
According to Fig. 13, the outside of the rigid-flexible integral link OA is fixed with the torque sensor through torque snesor base
(undeformed configuration, Fig. 14a). The outer ring of roller B bearing is installed in the sliding groove of the inner and outer cams
and its inner ring is fixed on the slider C through the shaft (Fig. 14b). When the inner and outer cams move in reverse at the same speed,
the roller B moves along the guide in radial direction and the length of link OB can be changed; when the cams move synchronously,
the roller B moves in tangential direction with a constant length of link OB and the leaf branches can be deformed by the rollers D1~D3
on link AC (deformed configuration, Fig. 14c), which is same as that in Fig. 11. In Fig. 14c, a steel wire cable is used to pull the link OA
temporarily for the deformed configuration demonstration, with OA=OB. Then, the link OA can remain stationary. The short-dashed
line indicates the initial position of link OB. It is also the neutral plane of the middle leaf branch with no deformation. After loaded by
the motors, the link OB reaches the position of the dotted line. It can be clearly observed that the middle leaf branch deviates from its
neutral plane. At same time, the link AC (long-dashed) rotates around the joint A, with an angle of α=0.22 rad. It is the deflection angle
of the equivalent torsion spring designed. The difference between the short-dashed line and dotted line is the deflection angle of the
stiffness adjustment module, with θ=0.44 rad. For specific adjustment principles, please refer to the paper [54]. The inner and outer
cams are drived by two independent motors with magnetic encoder (MBR320, 3200 pulses per revolution). The other end of the torque
sensor(JNNT-F-30Nm, the accuracy is 0.1 Nm) is fixed with frame. The motors housing are also fixed with the frame, and the freedom
of the motor comes from the leaf deformation. The overall structure of the measurement platform is shown in Fig. 14d. The mea­
surement process are as follows:

(1) Through the differential movement of the cam group, the roller B is driven to move along the link AC until OA=OB;

Fig. 13. Principle of torque-deflection test.

13
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

Fig. 14. Torsion spring measurement platform.

(2) Drive the cam group synchronously, and make the leaf to deform through the bearings on the link AC. Collect the information of
torque sensor and motor encoder. Motor encoder value can be converted to angle θ, which is − 0.44rad~0.44 rad, and the step
distance is 0.01 rad.

Similarly, the measurement data are fitted with a fifth-order polynomial odd function. The measurement curve and fitting curve are
as Fig. 15 shows. Subscript m represents the polynomial coefficients directly fitted according to the data measured through the stiffness
adjustment module and subscript n represents the polynomial coefficients equivalent replaced to the torsion spring. The equivalent
replacement relationship is Bn1 = 4Bm1, Bn2 = 16Bm2, Bn3 = 64Bm3.
According to the fifth-order fitting curve equation of the measurement result, Bm1 = 21.29, Bm2 = − 11.27, Bm3 = 1.84, Bn1 = 4Bm1
= 85.16, Bn2 = 16Bm2 = − 180.32, Bn3 = 64Bm3 = 117.76. In a likewise fashion, the torsion spring stiffness ksm and torque τsm are fitted
as:
{
τsn = − 180.32α2 + 85.16α
(29)
ksn = − 540.9α2 + 85.16

The constant item of the ksn is 85.16, which is 18% smaller than the FEM result. When α reaches 0.22 rad, the stiffness ksn is attenuated
from 85.16–58.98Nm/rad, and the attenuation rate is 30.74%. The impact of high-order items is remarkable and cannot be ignored.

4.4. Quantitative analysis of experiment error

Effect of material Young’s modulus E. When α→0 (as situation 1), the influences of the item cos(ε+α) and the material stress-strain
nonlinear item on the torsion spring stiffness ks are ruled out. As there is ks∝E according to Eq. (14), the key difference between the FEM
mode and prototype in small deformation mainly lies in its Young’s modulus E difference due to raw material reasons. For α→0, the
actual Young’s modulus E’ is:
limksn (α)
E = α→0

E = 169MPa
limksa (α)
α→0

The simulation data are normalized as:


′/
(30)

τsa = τsa E E

The τ’sa represents torque-deformation relationship after the Young’s modulus in the simulation model is changed to 169 MPa, as
Fig. 16 shows. After normalization, the Young’s modulus of the simulation model is the same as that of the actual prototype. Although
the force-deformation curves of the τ’sa and τsn are very close, there is still deviation.
Influence of the geometrical nonlinearity. The proposed analytical model assumes that the large deformation is not obvious with a

14
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

Fig. 15. Measurement and fitting results of torque-deflection relationship on the stiffness adjustment module.

small deflection constrain. When α reaches 0.22 rad, the actual deformation rate on the middle leaf branch is less than 6.67% and the
rate on the side leaf branch is less than 6.64%. In this case, the influence of the large deformation can be less than 0.5% after re-
simulation as Figs. 8 and 9 with a different load force. So, we are optimistic that the influence of the linear assumption exists, but
can be neglected.
Effect of the material behavior nonlinearity. In the FEM mode, the material behavior is ideal linear deformation (as situation 2). But in
the actual prototype, it is not so (as situation 3). As the difference in Young’s modulus has been eliminated before, the deviation
between τsn and τ’sa caused by the non-linearity of the material behavior can be expressed as:
( )
a w3 a w5
τsn − τsa = 1 + 2 (31)

cos(ε(w) + α(w))DR
3 5

The item cos(ε(w)+α( w)) can be decoupled by Eq. (32), and the non-linearity of the leaf material behavior can be evaluated by:

τsn − τ sa a1 b2 α2 a2 b4 α4
n(α) = = + (32)
τ′ sa 3a0 5a0

Substitute Eqs. (29) and (30) into the formula (32), and the result is shown in Fig. 17. Since the maximum value is less than 4.5%, it can
be considered that the leaf material does not reach the point σp on the stress-strain curve, and there is only weak nonlinear phe­
nomenon of the leaf material behavior. Another reason for the above deviation is the test error. Although it is difficult to quantitatively
distinguish the material nonlinear, geometrical nonlinearity, and test error as the error is less than 4.5%, this result can strongly prove
the correctness of the prototype design.

5. Conclusions

In order to tackle the contradiction of the torsion spring between the deflection and load in a compact way, this paper presents a
novel planar high-compliance joint. Its high compactness stems from its full use of its surrounding parts. Torsion spring is not an
additional independent part but a synergistic result of two adjacent parts developed from the stiffness adjustment linkage mechanism.
The flexible element is a cantilever leaf feature on the developed rigid-flexible integral linkage, which allows easily operation mode
conversion from the leaf to torsion spring. The deflection and load ability of torsion spring is enhanced by uniform stress design and
multiple leaf branches in the radial direction around the compliant hinged joint. Explicit analytical expression is established in a form

Fig. 16. Comparative analysis of simulation and measurement results.

15
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

Fig. 17. Evaluation result of leaf material behavior nonlinear.

convenient for leaf design and error analysis, and verified by the torque-deflection relationship of the finite element mode and pro­
totype. Error quantitative analysis shows that the difference of the material Young’s modulus is the main error factor in nearly zero
deformation. The geometric nonlinearity and material behavior nonlinearity are both very weak, and have little effect on the design
results. The influence of the Young’s modulus error can be compensated by a simple adjustment of the leaf thickness coefficient in the
future work.

CRediT authorship contribution statement

Fanghua Mei: Conceptualization, Methodology, Investigation, Visualization, Methodology, Writing – original draft. Shusheng Bi:
Conceptualization, Writing – review & editing, Supervision. Linkun Chen: Formal analysis. Hanjun Gao: Conceptualization, Funding
acquisition, Writing – review & editing.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper. Authors state no conflict of interest.

Acknowledgment

This work was supported by the National Natural Science Foundation of China (Grant Nos. 52075013 and 51675015).

References

[1] N. Govindan, S. Ramesh, A Thondiyath, Design of a variable stiffness joint module to quickly change the stiffness and to reduce the power consumption, IEEE
Access 8 (2020) 138318–138330.
[2] M. Malosio, G. Spagnuolo, A. Prini, et al., Principle of operation of RotWWC-VSA, a multi-turn rotational variable stiffness actuator, Mech. Mach. Theory 116
(2017) 34–49.
[3] B. Vanderborght, A. Albu-Schaeffer, A. Bicchi, et al., Variable impedance actuators: a review, Rob. Auton. Syst. 61 (12) (2013) 1601–1614.
[4] S. Wolf, G. Grioli, O. Eiberger, et al., Variable stiffness actuators: review on design and components, IEEE ASME Trans. Mechatron. 21 (5) (2015) 2418–2430.
[5] D. Gomez-Vargas, D. Casas-Bocanegra, M. Múnera, et al., Variable stiffness actuators for wearable applications in gait rehabilitation. Interfacing Humans and
Robots for Gait Assistance and Rehabilitation, Springer, Cham, 2022, pp. 193–212.
[6] L. Liu, S. Leonhardt, C. Ngo, et al., Impedance-controlled variable stiffness actuator for lower limb robot applications, IEEE Trans. Autom. Sci. Eng. 17 (2) (2019)
991–1004.
[7] S. Mghames, M. Laghi, C. Della Santina, et al., Design, control, and validation of the variable stiffness exoskeleton flexo, in: Proceedings of the 2017
International Conference on Rehabilitation Robotics (ICORR), IEEE, 2017, pp. 539–546.
[8] Y. Ning, Y. Liu, F. Xi, et al., Human-robot interaction control for robot driven by variable stiffness actuator with force self-sensing, IEEE Access 9 (2020)
6696–6705.
[9] B. Shusheng, L. Chang, Z. Xiaodong, et al., Review of structural research on adjustable stiffness actuators, J. Mech. Eng. 54 (13) (2018) 34–46.
[10] N.G. Tsagarakis, M. Laffranchi, B. Vanderborght, et al., A compact soft actuator unit for small scale human friendly robots, in: Proceedings of the 2009 IEEE
International Conference on Robotics and Automation, Kobe, IEEE, 2009, pp. 4356–4362.
[11] C.P. Vo, V.D. Phan, T.H. Nguyen, et al., A Compact adjustable stiffness rotary actuator based on linear springs: working principle, design, and experimental
verification, Actuators 9 (4) (2020) 141.
[12] S. Wolf, O. Eiberger, G. Hirzinger, The DLR FSJ: energy based design of a variable stiffness joint, in: Proceedings of the 2011 IEEE International Conference on
Robotics and Automation, Shanghai, IEEE, 2011, pp. 5082–5089.
[13] C.B. Yigit, E. Bayraktar, P. Boyraz, Low-cost variable stiffness joint design using translational variable radius pulleys, Mech. Mach. Theory 130 (2018) 203–219.
[14] J. Guo, Conceptual mechanical design of antagonistic variable stiffness joint based on equivalent quadratic torsion spring, Sci. Prog. 103 (3) (2020) 1–49,
003685042094129.
[15] Y. Ayoubi, M.A. Laribi, S. Zeghloul, et al., V2SOM: a novel safety mechanism dedicated to a Cobot’s rotary joints, Robotics 8 (1) (2019) 18.
[16] J. Liang, Y. Ning, H. Huang, et al., Design and modelling of a modular variable stiffness actuator, in: Proceedings of the 2019 IEEE International Conference on
Cyborg and Bionic Systems (CBS), IEEE, 2019, pp. 147–152.

16
F. Mei et al. Mechanism and Machine Theory 174 (2022) 104889

[17] R. Mengacci, M. Garabini, G. Grioli, et al., Overcoming the torque/stiffness range tradeoff in antagonistic variable stiffness actuators, IEEE ASME Trans.
Mechatron. 26 (6) (2021) 3186–3197.
[18] A. Jafari, N.G. Tsagarakis, I. Sardellitti, et al., A new actuator with adjustable stiffness based on a variable ratio lever mechanism, IEEE ASME Trans. Mechatron.
19 (1) (2014) 55–63.
[19] H. Jin, D. Yang, Z. Hui, et al., Flexible actuator with variable stiffness and its decoupling control algorithm: principle prototype design and experimental
verification, IEEE ASME Trans. Mechatron. 23 (3) (2018) 1279–1291.
[20] Y. Liu, X. Liu, Z. Yuan, et al., Design and analysis of spring parallel variable stiffness actuator based on antagonistic principle, Mech. Mach. Theory 140 (2019)
44–58.
[21] Y. Zhu, Q. Wu, B. Chen, et al., Design and evaluation of a novel torque-controllable variable stiffness actuator with reconfigurability, IEEE ASME Trans.
Mechatron. 27 (2021) 292–303.
[22] Z. Li, S. Bai, O. Madsen, et al., Design, modeling and testing of a compact variable stiffness mechanism for exoskeletons, Mech. Mach. Theory 151 (2020),
103905.
[23] Y. Ning, H. Huang, W. Xu, et al., Design and implementation of a novel variable stiffness actuator with cam-based relocation mechanism, J. Mech. Robot. 13 (2)
(2020) 1–22.
[24] R. Furnémont, G. Mathijssen, T. Van Der Hoeven, et al., Torsion MACCEPA: a novel compact compliant actuator designed around the drive axis, in: Proceedings
of the 2015 IEEE International Conference on Robotics and Automation (ICRA), IEEE, 2015, pp. 232–237.
[25] Y. Shao, W. Zhang, X. Ding, Configuration synthesis of variable stiffness mechanisms based on guide-bar mechanisms with length-adjustable links, Mech. Mach.
Theory 156 (2021), 104153.
[26] F. Negrello, M.G. Catalano, M. Garabini, et al., Design and characterization of a novel high-compliance spring for robots with soft joints, in: Proceedings of the
2017 IEEE International Conference on Advanced Intelligent Mechatronics (AIM), IEEE, 2017, pp. 271–278.
[27] Y. Xu, K. Guo, J. Li, et al., A novel rotational actuator with variable stiffness using S-shaped springs, IEEE ASME Trans. Mechatron. 26 (4) (2020) 2249–2260.
[28] M. Arefi, G.H Rahimi, Nonlinear analysis of a functionally graded beam with variable thickness, Sci. Res. Essays 8 (6) (2013) 256–264.
[29] P. Liu, G. Yao, L. Yan, Design of a helix-based revolute flexure hinge with large stroke, in: Proceedings of the 2021 IEEE International Conference on
Manipulation, Manufacturing and Measurement on the Nanoscale (3M-NANO), IEEE, 2021, pp. 117–121.
[30] Y. Chen, M. Yang, D. Wu, et al., Static modeling of a cross-spring flexure pivot with variable cross-section, in: Proceedings of the International Symposium on
Flexible Automation 2018 International Symposium on Flexible Automation, The Institute of Systems, Control and Information Engineers, 2018, pp. 25–29.
[31] H. Zhao, S. Bi, Stiffness and stress characteristics of the generalized cross-spring pivot, Mech. Mach. Theory 45 (3) (2010) 378–391.
[32] R.M. Panas, F. Sun, L. Bekker, et al., Combining cross-pivot flexures to generate improved kinematically equivalent flexure systems, Precis. Eng. 72 (2021)
237–249.
[33] Y. Miao, Z. Du, D. Wei, et al., Design and modeling of a variable thickness flexure pivot, J. Mech. Robot. 11 (1) (2018) 014502.
[34] Y. Xu, K. Guo, J. Sun, et al., Design, modeling, and control of a reconfigurable variable stiffness actuator, Mech. Syst. Signal Process. 160 (2) (2021), 107883.
[35] S.S. Bi, et al., Design and analysis of a novel variable stiffness actuator based on parallel-assembled-folded serial leaf springs, Adv. Robot. 31 (18) (2017)
990–1001.
[36] J. Choi, S. Hong, W. Lee, et al., A robot joint with variable stiffness using leaf springs, IEEE Trans. Robot. 27 (2) (2011) 229–238.
[37] R. Furnemont, G. Mathijssen, T.V.D. Hoeven, et al., Torsion MACCEPA: a novel compact compliant actuator designed around the drive axis, in: Proceedings of
the IEEE International Conference on Robotics and Automation, 2015.
[38] Y. Li, S. Bi, C. Zhao, Analytical modeling and analysis of rhombus-type amplifier based on beam flexures, Mech. Mach. Theory 139 (2019) 195–211.
[39] S. Bi, Y. Li, H. Zhao, Fatigue analysis and experiment of leaf-spring pivots for high precision flexural static balancing instruments, Precis. Eng. 55 (2019)
408–416.
[40] H. Liu, D. Zhu, J Xiao, Conceptual design and parameter optimization of a variable stiffness mechanism for producing constant output forces, Mech. Mach.
Theory 154 (2020) 104033.
[41] P. Bilancia, G. Berselli, S. Magleby, L. Howell, On the modeling of a contact-aided cross-axis flexural pivot, Mech. Mach. Theory 143 (2020)
103618.1–103618.14.
[42] P.P. Valentini, M. Cirelli, E. Pennestrì, Second-order approximation pseudo-rigid model of flexure hinge with parabolic variable thickness, Mech. Mach. Theory
136 (2019) 178–189.
[43] J. Dearden, C. Grames, J. Orr, et al., Cylindrical cross-axis flexural pivots, Precis. Eng. 51 (2018) 604–613.
[44] S. Linß, P. Gräser, T. Räder, et al., Influence of geometric scaling on the elasto-kinematic properties of flexure hinges and compliant mechanisms, Mech. Mach.
Theory 125 (2018) 220–239.
[45] Z. Xie, L. Qiu, D. Yang, Analysis of a novel variable stiffness filleted leaf hinge, Mech. Mach. Theory 144 (2020), 103673.
[46] B.D. Jensen, L.L Howell, The modeling of cross-axis flexural pivots, Mech. Mach. Theory 37 (5) (2002) 461–476.
[47] N.P. Belfiore, M. Balucani, R. Crescenzi, M. Verotti, Performance analysis of compliant MEMS parallel robots through pseudo-rigid-body model synthesis, in:, in:
Proceedings of the ASME 11th Biennial Conference on Engineering Systems Design and Analysis, ESDA 3, 2012, pp. 329–334.
[48] S. Šalinic´, A. Nikolic´, A new pseudo-rigid-body model approach for modeling the quasi-static response of planar flexure-hinge mechanisms, Mech. Mach. Theory
124 (2018) 150–161.
[49] P.P. Valentini, E. Pennestrì, Compliant four-bar linkage synthesis with second-order flexure hinge approximation, Mech. Mach. Theory 128 (2018) 225–233.
[50] G. Palli, G. Berselli, C. Melchiorri, et al., Design of a variable stiffness actuator based on flexures, J. Mech. Robot. 3 (3) (2011) 34501.
[51] P. Bilancia, G. Berselli, G. Palli, Virtual and physical prototyping of a beam-based variable stiffness actuator for safe human-machine interaction, Robot.
Comput. Integr. Manuf. 65 (2020) 101886.
[52] Aiple M, Gregoor W, Schiele A. A Dynamic Robotic Actuator with Variable Physical Stiffness and Damping[J]. arXiv preprint arXiv:1906.07669, 2019.
[53] F. Sergi, D. Accoto, G. Carpino, et al., Design and characterization of a compact rotary series elastic actuator for knee assistance during overground walking, in:
Proceedings of the IEEE Ras & Embs International Conference on Biomedical Robotics & Biomechatronics, IEEE, 2012.
[54] F. Mei, S. Bi, C. Liu, et al., Optimal design of cam curve dedicated to improving load uniformity of bidirectional antagonistic VSA, in: Proceedings of the
International Conference on Intelligent Robotics and Applications, Cham, Springer, 2021, pp. 3–13.

17

You might also like