You are on page 1of 12

Elastic Averaging in Flexure

Mechanisms: A Three-Beam
Parallelogram Flexure Case
Study
Shorya Awtar1 Redundant constraints are generally avoided in mechanism design because they can lead
e-mail: awtar@umich.edu
to binding or loss in expected mobility. However, in certain distributed-compliance flex-
ure mechanism geometries, this problem is mitigated by the phenomenon of elastic aver-
Kevin Shimotsu aging. Elastic averaging is a design paradigm that, in contrast with exact constraint
design principles, makes deliberate and effective use of redundant constraints to improve
Shiladitya Sen performance and robustness. The principle of elastic averaging and its advantages are
illustrated in this paper by means of a three-beam parallelogram flexure mechanism,
Mechanical Engineering,
which represents an overconstrained geometry. In a lumped-compliance configuration,
University of Michigan,
this mechanism is prone to binding in the presence of nominal manufacturing and as-
2350 Hayward Street,
sembly errors. However, with an increasing degree of distributed-compliance, the mecha-
Ann Arbor, MI 48109
nism is shown to become more tolerant to such geometric imperfections. The nonlinear
elastokinematic effect in the constituent beams is shown to play an important role in
analytically predicting the consequences of overconstraint and provides a mathematical
basis for elastic averaging. A generalized beam constraint model is used for these pre-
dictions so that varying degrees of distributed compliance are captured using a single
geometric parameter. The closed-form analytical results are validated against finite ele-
ment analysis, as well as experimental measurements. 关DOI: 10.1115/1.4002204兴

Keywords: elastic averaging, constraint-based design, overconstraint, flexure


mechanisms, beam constraint model, elastokinematic effect

1 Introduction and Motivation “perfectly” parallel. Otherwise, the DoF count will drop to zero,
as formally predicted by Grubler’s criterion 关14兴, which does not
Exact constraint design 共ECD兲, also known as kinematic de-
account for special geometries.
sign, has been the cornerstone of rigid-link mechanism analysis
This three-link parallelogram mechanism 共Fig. 1共a兲兲 is clearly
and synthesis 关1–13兴. The key guiding principle in this paradigm
overconstrained. The distinctive feature of this arrangement is that
is to employ exactly the minimum number of constraints that are
needed to produce the desired mobility or degrees of freedom its mobility is critically dependent on its geometric accuracy, mak-
共DoFs兲. ECD ensures predictable and repeatable motions, elimi- ing it highly sensitive to manufacturing tolerances and environ-
nates any possibility of binding, and allows cost-effective fabrica- mental variations. The slightest loss of parallelism between the
tion of mating components. In contrast, mechanism geometries three links would lead to a geometry that can move only if small
that employ redundant constraints, also known as overconstrained displacements along the length of the links were allowed. The
geometries, are often prone to undesired internal loads and bind- latter is impossible given the infinite translational stiffness of the
ing. rigid links and the revolute joints. This would lead to binding or a
The rationale for avoiding redundant constraints, even when loss in mobility, as predicted above by Grubler’s criterion. Thus,
they are apparently nonconflicting, lies in the ideal constraint overconstrained rigid-link arrangements are feasible only when
characteristics of traditional rigid-link mechanisms: The links are the manufacturing accuracy is extremely high, or if the revolute
perfectly rigid, the joints are infinitely stiff in their constraint di- joints are loose enough to accommodate any residual manufactur-
rections, while offering zero resistance in their DoF directions, ing errors.
and the joints are free of looseness or backlash. These idealiza- Given this common-sense rationale for ECD, it has also been
tions are indeed closely maintained even in the physical embodi- extended to the design of flexure mechanisms 共also known as
ments of rigid-link mechanisms. The parallelogram mechanism of compliant mechanism兲, which comprise rigid and compliant ele-
Fig. 1共a兲, without the middle link, would be exactly constrained ments 关12,13,15兴. Motion in a flexure mechanism arises from the
with the two parallel rigid links, each connecting the ground to the elastic deformation of the compliant elements instead of sliding or
coupler via two revolute joints. One way to describe the mobility rolling 关11–13,16,17兴. The resulting lack of friction and backlash
of this mechanism is to say that the coupler has one translational provides smooth motion and an exceptional level of precision.
DoF along the transverse direction, as shown in this figure. When The monolithic construction of flexure mechanisms also leads to
a third rigid link 共shown dashed兲 is added between the ground and design simplicity and maintenance-free operation even in harsh
coupler, this one DoF is preserved only if the three links are environments. In spite of these benefits, flexure mechanisms ex-
hibit several performance trade-offs that arise from their nonideal
constraint behavior 关17,18兴. Parallelogram flexure mechanisms
1
Corresponding author. with a single translational DoF are shown in shown in Fig. 1共b兲
Contributed by the Mechanisms and Robotics Committee of ASME for publica-
tion in the JOURNAL OF MECHANISMS AND ROBOTICS. Manuscript received November
共lumped compliance兲 and Fig. 1共c兲 共distributed compliance兲.
12, 2009; final manuscript received July 6, 2010; published online September 30, These are commonly used as linear guides or bearings to provide
2010. Assoc. Editor: G. K. Ananthasuresh. approximate straight line motion in various applications

Journal of Mechanisms and Robotics Copyright © 2010 by ASME NOVEMBER 2010, Vol. 2 / 041006-1
Coupler Y fy
Transverse mz
z
a.
Z X fx
Axial uy

b b ux
1

Ground Fig. 2 Generalized beam flexure

Coupler Coupler

b. c. largely empirical or qualitative 关26–30兴. To enable a wider adop-


tion of elastic averaging, this paper provides a systematic and
mathematical basis that helps answer some key design questions
such as the following: When is elastic averaging applicable in
flexure mechanism design? How large geometric imperfections
can be tolerated in an overconstrained flexure mechanism without
Ground risking binding? How does one quantify the presence and benefits
of elastic averaging?
Fig. 1 Parallelogram mechanism: „a… traditional linkage, „b… Such questions are addressed in this paper in the context of the
lumped-compliance flexure, and „c… distributed-compliance three-beam parallelogram flexure mechanism. Section 2 provides
flexure a generalized parametric beam constraint model that captures the
constraint characteristics 共stiffness and error motion兲 of a variable
cross-section beam flexure. This model accommodates varying
关11–13,16–18兴. Following ECD guidelines, let us initially con- degrees of distributed compliance via a single geometric param-
sider these mechanisms with only two beams each. For these eter and captures relevant geometric nonlinearities. In Sec. 3, the
mechanisms to serve as motion guides, it is generally desirable to generalized beam constraint model is used to develop the closed-
maximize their load-bearing stiffness along the axial direction, form load-displacement relations for a three-beam parallelogram
while keeping the transverse stiffness as low as possible to mini- flexure with a nominal geometric imperfection. In addition to con-
mize stresses and maximize range along their single translational firming the qualitative arguments behind elastic averaging, these
DoF. While the lumped-compliance configuration provides better closed-form analytical results provide new physical and quantita-
axial stiffness, its transverse stiffness is also relatively high. This tive insight into overconstraint in flexure mechanisms. The design
leads to greater actuation effort, higher stresses, and reduced range of a novel, reconfigurable experimental set-up to test various geo-
of motion along the DoF. DoF motion range is improved in the metric configurations of the three-beam parallelogram flexure is
distributed-compliance configuration, which provides a better dis- presented in Sec. 4. The closed-form analytical results are vali-
tribution of strains and therefore lower stresses. But now the axial dated via comparison with finite element analysis and experimen-
stiffness is also reduced, which limits the load-bearing capacity. tal measurements in Sec. 5. The contributions of this paper are
Any attempt to increase this axial stiffness by increasing the flex- summarized in Sec. 6.
ure beam thickness results in a cubic order increase in the trans-
verse stiffness, which is obviously undesirable. These perfor-
mance trade-offs seen in the lumped and distributed-compliance 2 Background: Beam Constraint Model for a Variable
geometries are inherent to flexure mechanisms 关18兴. Cross Section Beam
Contradicting ECD guidelines, if one were to introduce a “per-
fectly” parallel third beam 共dashed line兲 in either of the parallelo- To study elastic averaging in the three-beam parallelogram flex-
gram flexure mechanisms of Fig. 1共b兲 or Fig. 1共c兲, there would be ure mechanisms of Figs. 1共b兲 and 1共c兲, it is first important to
a proportional increase in the transverse and axial stiffness values. establish the constraint properties of the beam flexure. This has
Thus, the introduction of additional beams clearly helps overcome been previously accomplished via a beam constraint model
the above performance trade-offs. But the question that remains is 关18,31兴, a brief summary of which is provided here. Figure 2
whether the problems associated with overconstraint, described illustrates an initially straight, variable cross section beam in its
earlier, are applicable here or not. It may be qualitatively argued undeformed and deformed states with normalized end-loads 共fx,
that overconstraint will likely be a problem for the lumped- fy, and mz兲 and end-displacements 共ux, uy, and ␪z兲 along the X-Y-Z
compliance configuration 共Fig. 1共b兲兲 since the flexural pivots pre- coordinate frame. Displacements and lengths are normalized by
clude the possibility of “loose” joints and also provide relatively the overall beam length L, forces by EIzz / L2, and moments and
high translational stiffness. Any deviation from perfect parallelism strain energy by Izz / L. All normalized quantities are scalar and are
can still lead to an unexpectedly high stiffness and potential bind- represented by lower case letters throughout this paper; addition-
ing in the transverse or DoF direction. However, in the three-beam ally, all loads 共forces/moments兲 are represented by bold letters. E
distributed-compliance configuration of Fig. 1共c兲, the axial stiff- represents Young’s modulus of the material for an XY plane-stress
ness of each individual beam is relatively lower and therefore the condition and the plate modulus for an XY plane-strain condition.
resulting increase in the transverse stiffness of the mechanism, in Izz represents the second moment of area of the two compliant
case of geometric imperfections, is also relatively small. Conse- end-segments, each of length b, thickness t, and depth h. The
quently, binding or mobility loss may be avoided. It is therefore middle section of the beam is thick enough to be considered rigid.
no surprise that distributed-compliance multibeam parallelogram The geometric parameter b quantifies the degree of distributed
flexure mechanism design has indeed been employed in the past compliance: b = 1 / 2 represents a simple beam with uniformly dis-
关19–21兴. tributed compliance, while b → 0 corresponds to a lumped-
This ability of distributed-compliance topologies to tolerate compliance configuration.
overconstrained geometric arrangements without binding, even in Clearly, the transverse directions Y and ⌰ represent DoFs,
the presence of small geometric imperfections, is known as elastic while the axial direction X represents a degree of constraint
averaging 关8–12,21–25兴. While elastic averaging is not a new idea 共DoC兲. The beam constraint model 共BCM兲 comprises parametric
and has been exploited occasionally to achieve greater perfor- nonlinear load-displacement and strain energy relations for the
mance in mechanisms and machines, its treatment so far has been beam:

041006-2 / Vol. 2, NOVEMBER 2010 Transactions of the ASME


再冎冋fy
mz
=
共0兲
k11
共0兲
k12
共0兲
k12
共0兲 册再 冎 冋uy
␪z
+ fx
共1兲
k11
共1兲
共1兲
k12
共1兲 册再 冎 冋
uy
␪z
+ f2x
共2兲
k11
共2兲
共2兲
k12
共2兲 册 Table 1 Beam characteristic coefficients for a simple beam

再冎
k22 k12 k22 k12 k22
k共0兲
11 12
uy
⫻ 共1兲 k共0兲
12 ⫺6
␪z
k共0兲

冋 册再 冎
22 4
共1兲 共1兲
fx 1 k11 k12 uy k共1兲 6/5
ux = − 兵uy ␪z 其 共1兲 共1兲
11
k33 2 k12 k22 ␪z

册再 冎
k共1兲 ⫺1/10


12
共2兲 共2兲
k11 k12 uy k共1兲 2/15
− fx兵uy ␪z 其 共2兲 共2兲 共2兲 22
k12 k22 ␪z
k共2兲 ⫺1/700

冋 册再 冎
11
共0兲 共0兲
1 f2x
1 k11 k12 uy k共2兲 1/1400
+ 兵uy ␪z 其
12
v= 共0兲 共0兲
2 k33 2 k12 k22 ␪z k共2兲 ⫺11/6300

冋 册再 冎
22
共2兲 共2兲
1 k11 k12 uy
− f2x兵uy ␪z 其 共2兲 共2兲 共3兲
2 k12 k22 ␪z
The BCM is based on the Euler–Bernoulli and linearized beam- These coefficients are plotted in Fig. 3 共for t = 0.002413兲 to
curvature assumptions. Furthermore, the application of load equi- illustrate their dependence on the degree of distributed compliance
librium in the deformed beam state leads to the presence of the 共b兲. Such a graphical visualization proves to be helpful in inter-
axial load fx in the transverse load-displacement relation 共1兲, the preting the results associated with elastic averaging in Sec. 3.
geometric constraint relation 共2兲, and the strain energy relation Overall, relations 共1兲–共4兲 provide a single model for beams with
共3兲. All three relations initially exist as infinite series in the axial any desired degree of distributed compliance, captured simply via
load fx but are truncated in a mutually consistent fashion to retain the geometric variable b.
only the relevant powers of fx in the BCM 关32兴.
In relation 共1兲, the first matrix captures elastic stiffness, while 3 Three-Beam Parallelogram Flexure Mechanism
the second matrix captures load-stiffening, which quantifies the A three-beam parallelogram flexure mechanism comprising
change in DoF direction stiffness in the presence of a constraint ground, motion stage, and three “nominally” parallel beams is
load. The third matrix is included to maintain consistency with the illustrated in Fig. 4. Although simple beams are shown for clarity,
following two relations. In relation 共2兲, the first term represents this discussion and analytical treatment are equally valid for the
the elastic stretching of the beam. The second term, which repre- generalized beam discussed in Sec. 2. While beams 1 and 2 are
sents a kinematic component, arises from the geometric constraint assumed perfectly parallel, a parallelism error ␣ is introduced at
of constant beam arc length and is exclusively dependent on the
beam 3. Angle ␣ simulates a typical manufacturing or assembly
transverse displacements uy and ␪z. The third term, also arising imperfection, for example, a 1 mm parallelism error over 100 mm
from beam arc-length conservation, depends on the transverse dis- beam length. This provides an estimate of the maximum magni-
placements, as well as the axial load, and is therefore referred to
tude of ␣ that might be of interest: 0.01. Consequently, the small
as the elastokinematic component. This term is unique to
angle approximations cos ␣ ⬇ 1 and sin ␣ ⬇ ␣ are applicable. The
distributed-compliance configurations and results in an additional
normalized loads fx, fy, and mz applied at point O on the motion
compliance along the constraint direction that increases quadrati-
stage result in normalized displacements x, y, and ␪ at the same
cally with the DoF displacements. Relation 共3兲 presents a nonlin-
point. The normalization scheme followed here is the same as
ear strain energy expression for the beam that is compatible with
described above.
the first two relations.
As discussed in Sec. 1, the presence of three beams makes this
The nondimensional coefficients k in relation 共1兲 are function-
mechanism geometry overconstrained. If the beams were all per-
als of the beam shape and are referred to as the beam character-
fectly parallel, one would expect 1 DoF 共Y direction displace-
istic coefficients. These coefficients are identified via a subscript
that highlights their respective location in the stiffness matrix that ment兲 and 2 DoCs 共X direction displacement and ⌰ direction ro-
relates the transverse loads and displacements, and a superscript tation兲. However, if beam 3 共or any other, for that matter兲 is
that denotes the power of axial load f x1 in the transverse load-
displacement relation that the coefficient are associated with. It is 7 −4
noteworthy that the same coefficients repeat in relations 共2兲 and 250 2.5
x 10 x 10
7.5
共3兲, which highlights the fact that these three relations are funda- (2)
mentally inter-related 关32兴. For the limiting case of a simple beam − k11
Elastokinematic Coefficient (− k )
(2)

共uniform thickness and initially straight兲 given by b = 0.5, these


22

200 2 6
Axial Elastic Stiffness (k33 )

beam characteristic coefficients assume the following numerical


(0)
Elastic Stiffness (k11 )

values listed in Table 1.


For a generalized beam shape with b ⬍ 0.5, these characteristics 150 1.5 4.5

coefficients have been analytically derived previously 关18兴. Only


those that are relevant for the subsequent discussion is this paper
are listed below: 100 1 3
k
33

6 k(1)
共0兲 11
k11 =
b共3 − 6b + 4b2兲 50 0.5 1.5

共2兲 − 2b 共105 − 630b + 1440b − 1480b + 576b 兲


3 2 3 4
k11 =
175共3 − 6b + 4b2兲3 0 0
0 0.1 0.2 0.3 0.4
0
0.5

冉 冊
Beam Shape Parameter (b)
1 12
k33 = 共4兲
2b t2 Fig. 3 Beam characteristic coefficients versus beam shape

Journal of Mechanisms and Robotics NOVEMBER 2010, Vol. 2 / 041006-3


(ux(1), uy(1))

C1 Beam 1

fy
Y Beam 3
(ux(3), uy(3)) mz w
(x, y) fx

Z
X O

w
(ux(2), uy(2))
Ground Beam 2

Motion C2
Stage
1

Fig. 4 Three-beam parallelogram flexure with a geometric error

slightly angled, then a potential loss in mobility can be physically virtual work is employed here to avoid the internal loads alto-
visualized. In the present case, the arrangement of beam 1, beam gether and considerably reduce the mathematical complexity 关32兴.
3, and the motion stage provides a remote center of rotation at C1 Since the shape of each beam is considered identical, the strain
for the motion stage 关18兴; on the other hand, beam 2, beam 3, and energy associated with the ith beam may be determined by sub-
the motion stage result in a remote center of rotation located at C2. stituting the beam axial force obtained from Eq. 共2兲 into Eq. 共3兲.

冉 冋 册再 冎冊
Thus, upon the application of a Y direction force at the motion
共1兲 共1兲 u y 共i兲 2
stage, one part of the mechanism works to make the motion stage 1 k11 k12
rotate counterclockwise about point C1, while the other part tries u x共i兲 + 兵 u y 共i兲 ␪ z共i兲 其 共1兲 共1兲
1 2 k12 k22 ␪ z共i兲

冋 册再 冎冊
to make it rotate clockwise about C2. This obviously imposes vi = k33


conflicting geometric requirements on the motion stage. If the 2 共2兲 共2兲
k11 k12 u y 共i兲
constraint provided by each beam were close to ideal, i.e., high 1 − k33兵uy共i兲 ␪z共i兲 其 共2兲 共2兲
␪ z共i兲

册再 冎
stiffness in the beam’s respective constraint 共or axial兲 direction, k12 k22
then the Y DoF direction stiffness of the overall mechanism would
also increase, potentially leading to binding. The remainder of this
section provides a closed-form analytical corroboration of this
1 k11

共0兲 共0兲
k12
+ 兵 u y 共i兲 ␪ z共i兲 其 共0兲 共0兲
2 k12 k22
u y 共i兲
␪ z共i兲
共7兲
qualitative argument. The total strain energy is simply the sum of the strain energies
In the presence of beams 1 and 3 only, the motion stage rotation of the individual beams, which may be obtained by substituting
would be approximately given by y ␣ / w, whereas beams 2 and 3, Eqs. 共6兲 into Eq. 共7兲. The principle of virtual work then dictates
by themselves, would produce −y ␣ / w 关18兴. Therefore, it is rea-
sonable to assume that the actual motion stage rotation will be ␦ 共 v 1 + v 2 + v 3兲 = ␦ w = f x␦ x + f y␦ y + m z␦ ␪ 共8兲
bounded as follows:
Since x, y, and ␪ are independent displacement coordinates in
y␣ y␣ this problem, the coefficients of their variations in the above equa-
− ⬍␪⬍ 共5兲
w w tion may be identically set to zero to obtain load-displacement
relations for the X, Y, and ⌰ directions. The X direction relation is
Since we are generally interested in a DoF motion range within
given by
⫾0.1, the above inequality implies that the motion stage rotation
should be of the order of 0.001 or less. Thus, the small angle k33
approximations cos ␪ ⬇ 1 and sin ␪ ⬇ ␪ are also justified here.

册再 冎冊
fx =
With this knowledge, we proceed to state the conditions of geo-
metric compatibility between the three beams by expressing the
end-displacements of each beam 共measured along their local co-
冉 1 − k33兵y ␪ 其 冋 共2兲
k11
共2兲
k12
共2兲
k12
共2兲
k22
y

ordinate axes aligned with the undeformed state of the respective
beam兲 in terms of the motion stage displacements:
Beam 1: ux共1兲 ⬇ x − w␪, uy共1兲 = y − w共1 − cos ␪兲 ⬇ y, ␪ z共1兲 = ␪
⫻ 冋 3
2

共1兲
k11 共1兲
k12
兵y ␪ 其 共1兲 共1兲
k12 k22
册再 冎 y

+ 共3x − y ␣兲 册 共9兲

This result may be used to express the Y and ⌰ direction rela-


Beam 2: ux共2兲 ⬇ x + w␪, uy共2兲 = y + w共1 − cos ␪兲 ⬇ y, ␪ z共2兲 = ␪ tions in more concise forms:

Beam 3: ux共3兲 = x cos ␣ − y sin ␣ ⬇ x − y ␣, 共0兲


fy = 3共k11 共0兲
y + k12 共1兲
␪兲 + fx k11 冉
共1兲
y + k12 ␪−

3
1
冊 共2兲
+ f2x共k11
3
共2兲
y + k12 ␪兲
uy共3兲 = y cos ␣ + x sin ␣ ⬇ y − x␣, ␪ z共3兲 = ␪ 共6兲
k33 2
· y ␣2

册再 冎冊
+

冉 冋
A direct analysis of this problem would involve solving 12
共2兲 共2兲 3
equations simultaneously: the three constitutive load-displacement k11 k12 y
relations 共1兲 and 共2兲 for each beam and the three load equilibrium 1 − k33兵y ␪ 其 共2兲 共2兲
k12 k22 ␪
conditions for the motion stage. Assuming the three externally
applied loads 共fx, fy, and mz兲 to be known, the 12 unknowns in the
problem are the three internal end-loads for each beam and the
three motion stage displacements 共x, y, and ␪兲. Solving these 12 共2兲 共2兲
␪兲 ·
冉 2
共k33兲2 2␪2w2 + y 2␣2
3

册再 冎冊
+ 共k11 y + k12 共10兲
nonlinear equations is mathematically tedious and involves solv-
ing for the 9 internal end-loads, which are irrelevant to the final
results. Instead, an energy approach based on the principle of
冉 1 − k33兵y ␪ 其 冋 共2兲
k11
共2兲
k12
共2兲
k12
共2兲
k22
y

2

041006-4 / Vol. 2, NOVEMBER 2010 Transactions of the ASME


共0兲 共0兲 共1兲 共1兲 1 共2兲 共2兲
faster rate. Apart from its dependence on beam shape parameter b,
mz = 3共k12 y + k22 ␪兲 + fx共k12 y + k22 ␪兲 + f2x共k12 y + k22 ␪兲 k33 also increases quadratically2 with decreasing beam thickness t.
3
The elastokinematic coefficient k共2兲 11 , which is a unique attribute of
k33 distributed-compliance configurations, expectedly reduces with
· 2␪w2

册再 冎冊
+

冉 冋 共2兲 共2兲
decreasing values of b and, in fact, approaches zero asymptoti-
k11 k12 y cally for b less than 0.1. We would like to note here that the
1 − k33兵y ␪ 其 共2兲 共2兲
k12 k22 ␪ previously reported 关21兴 expression for the Y direction stiffness is

冉 冊 incorrect because of premature truncation. Since k33k共2兲 2


2 11 y is gen-
共k33兲2 2␪2w2 + y 2␣2 erally not small with respect to 1, an infinite series expansion and
共2兲 共2兲 3 truncation are not justifiable.
␪兲 ·

册再 冎冊
+ 共k11 y + k12

冉 冋 共2兲 共2兲 2
For nominal mechanism dimensions 共w = 0.5兲 and beam thick-
k11 k12 y ness 共t = 0.002413兲, a series of Y direction stiffness plots, based on
1 − k33兵y ␪ 其 共2兲 共2兲
k12 k22 ␪ Eq. 共14兲, over a range of beam shapes and imperfection angles are
共11兲 presented in Fig. 5. Based on these results, several important ob-
servations can be made:
Since ␪ is of the order of 0.001 or less for y of the order of 0.1, 1. In the absence of a geometric imperfection 共␣ = 0兲, the Y
the former may be dropped with respect to the latter in the above stiffness simply reduces to 3k共0兲
11 . The effect of beam shape param-
relations, incurring errors less than 1%. This simplifying assump- eter b on this nominal stiffness is captured in relation 共4兲.
tion reduces relation 共9兲 to 2. In the presence of small but finite ␣, a purely linear result
fx 1 f x 共2兲 2 y ␣ 共valid only for y → 0兲 may be obtained by dropping the higher
x⬇ − k共1兲y 2 − k11 y + 共12兲 power y terms in Eq. 共13兲:
3k33 2 11 3 3
fy ⬇ 共3k11
共0兲
+ 3 · k33␣2兲 y
2
In this expression, the first term may be identified to be the 共15兲
purely elastic component of the axial displacement, the second This shows that the axial stiffness k33 of the individual beams is
term represents the purely kinematic component, the third term reflected in the Y DoF direction due to the geometric imperfec-
represents the elastokinematic component, and the last term rep- tion. Consequently, the effective stiffness in the DoF direction is
resents a new kinematic term arising from the geometric imper- higher than the nominal stiffness and grows as a quadratic func-
fection. tion of ␣, as may be seen at y = 0 for various beam shapes in Figs.
Since we are interested primarily in the Y direction load- 5共a兲–5共c兲. With decreasing b, since k33 increases at a much faster
displacement characteristics, in addition to the above simplifica-
rate than k共0兲
11 共see Fig. 3兲, the increase in effective stiffness com-
tion we can also set mz = fx = 0 without loss in generality. This
pared with nominal stiffness is much more significant in the case
helps further simplify the mathematical relation 共10兲, so as to
of lumped compliance compared with disturbed compliance, for
allow some physical insight:
the same beam thickness t and geometric error ␣. For example,
共0兲
fy ⬇ 3k11 y+
2
· 冉k33
共2兲 2 · y ␣
3 1 − k33k11 y
2
冊 Fig. 5共d兲 shows that while the nominal stiffness increases by two
times going from a b value of 0.5 to 0.1, the effective stiffness

冉 冊
increases by more than four times. This increase in effective DoF
2
2 k33 共2兲 3 2 stiffness, seen at y = 0, is the underlying cause of binding in this
+ · 共2兲 2 · k11 y␣ 共13兲 overconstrained mechanism and is clearly more prominent in the
3 1 − k33k11 y
case of lumped compliance. Thus, the vulnerability of lumped
Differentiation with respect to y yields the DoF direction stiff- compliance to mobility loss in overconstrained mechanism geom-
ness of the three-beam parallelogram flexure with a small paral- etries is clear even in a purely linear analysis.
lelism error ␣, for any given degree of distributed compliance 3. Expressions 共13兲 and 共14兲 show that with increasing y dis-
captured via variable b: placement, the elastokinematic nonlinearity results in a reduced
⳵ fy
⳵y
共0兲
= 3k11
2
+ · 冉
k33
共2兲 2 · ␣
3 1 − k33k11 y
2
冊 constraint 共or axial兲 stiffness of the individual beams: k33 / 共1
− k33k共2兲
11 y 兲 instead of k33, which also leads to a reduction in the
2

DoF direction stiffness of the flexure mechanism. This happens


10
3
·冉 k33
共2兲 2
1 − k33k11 y
冊 2
共2兲 2 2
k11 y␣
when −k33k共2兲 11 y 共always a positive quantity兲 in the denominator
2

becomes of the order of or larger than 1.


For the case of thin beams 共high k33兲 with uniformly distributed
+
8
·冉 k33
共2兲 2
3 1 − k33k11 y
冊 3
共2兲 2 4 2
共k11 兲y␣ 共14兲
compliance, which corresponds to the maximum possible magni-
tude of k共2兲 共2兲 2
11 , the quantity −k33k11 y can become much larger than
1. This leads to the following simplified DoF load-displacement
Since the force and displacement values in this expression are
relation:
normalized with respect to EI / L2 and L, respectively, it presents a
nondimensional stiffness that is normalized with respect to EI / L3.
The beam characteristic coefficients that appear in this expression
are k共0兲 共2兲 共0兲
共0兲
fy ⬇ 3k11 y+
2
· 冉 k33
3 − k33k11 y
2

共2兲 2 · y ␣ +
2
·
k33
共2兲 2
3 − k33k11 y
冉 冊 2
共2兲 3 2
· k11 y␣
11 , k11 , and k33. Referring to Eq. 共1兲, k11 is the linear elastic 共0兲
Y-direction bending stiffness of an individual beam 共for zero or = 3k11 y
negligible ␪兲. Referring to Eq. 共2兲, k33 is the linear elastic This analytical observation is significant. It indicates that in
X-direction axial stiffness of an individual beam and k共2兲 11 repre- certain cases 共when −k33k共2兲
11 y Ⰷ 1兲, the effective stiffness of the
2
sents the elastokinematic effect in this direction associated with y three-beam parallelogram flexure with a geometric imperfection
displacement 共for zero or negligible ␪兲. The effective X-direction comes back down to its nominal stiffness as though there was no
compliance, resulting from the elastic and nonlinear elastokine-
matic effects, is given by 共1 / k33 − k共2兲
11 y 兲, and stiffness is simply
2
2
the inverse of this. Figure 3 illustrates that with increasingly This might appear counterintuitive since the axial beam stiffness increases lin-
early with beam thickness t. However, k33 represents the axial stiffness normalized
lumped-compliance, k共0兲 11 gradually grows to infinity. Similarly, the with respect to the bending stiffness, which leads to an inverse quadratic dependence
normalized axial stiffness k33 also grows to infinity, but at an even on t.

Journal of Mechanisms and Robotics NOVEMBER 2010, Vol. 2 / 041006-5


800 800
a. b = 0.5 c. α = 0.01 b = 0.1
700 700

600 600
Normalized DoF Stiffness

Normalized DoF Stiffness


500 500

400 400
α = 0.005
300 300

200 200
α = 0.01 α = 0.005
100 100
α=0
0 0
α=0
−100 −100
−0.15 −0.1 −0.05 0 0.05 0.1 0.15 −0.15 −0.1 −0.05 0 0.05 0.1 0.15
Normalized DoF Displacement (y) Normalized DoF Displacement (y)

800 800
b. b = 0.2 d. b = 0.1 α = 0.01
700 700

600 600

Normalized DoF Stiffness


Normalized DoF Stiffness

500 500
b = 0.2
400 400
α = 0.01
b = 0.3
300 300
α = 0.005 b = 0.5
200 200

100 100

α=0
0 0

−100 −100
−0.15 −0.1 −0.05 0 0.05 0.1 0.15 −0.15 −0.1 −0.05 0 0.05 0.1 0.15
Normalized DoF Displacement (y) Normalized DoF Displacement (y)

Fig. 5 Analytically predicted Y-direction stiffness of the three-beam parallelogram flexure

geometric imperfection. This is evident in the distributed- dicted to approach zero and even become negative. This poten-
compliance configuration in Fig. 5共a兲. In other words, because of tially leads to bistability, which is highly undesirable in a linear
the elastokinematic effect associated with distributed-compliance, bearing design. This simply reaffirms the importance and applica-
the −k33k共2兲
11 y Ⰷ 1 condition is met and the risk associated with
2 bility of ECD guidelines in lumped-compliance flexure mecha-
binding in the given overconstrained mechanism is mitigated. This nism geometries. On the other hand, the fact that such a dramatic
phenomenon is elastic averaging. stiffness variation and resulting bistability does not happen in the
On the other hand, −k33k共2兲 2
11 y can remain low despite increasing
corresponding distributed-compliance geometry highlights that
displacements 共y ⬃ 0.1兲 and thin beams 共i.e., high k33兲 if −k共2兲 11 is
ECD guidelines are not strictly applicable in this case. A compre-
small, as is the case with lumped-compliance configurations. As hensive prior-art summary and analytical treatment of multistabil-
per Fig. 3, even though k33 increases with decreasing b, the mag- ity in flexure mechanisms is covered in Ref. 关33兴.
4. Ultimately, this discussion highlights that to predict the na-
nitude of the elastokinematic coefficient k共2兲 11 drops and becomes ture and benefits of elastic averaging in flexure mechanisms, we
virtually zero for b ⬍ 0.1. Therefore, there is not much respite in
first have to recognize the elastokinematic nonlinearity in the axial
the constraint direction stiffness of the beams even with increasing
constraint stiffness of the constituent beam: k33 / 共1 − k33k共2兲11 y 兲.
2
y displacement, and k33 / 共1 − k33k共2兲
11 y 兲 remains close to k33. Thus,
2
Furthermore, for elastic averaging to take place, the condition
not only is the Y DoF effective stiffness in the presence of a
geometric imperfection higher in a lumped-compliance configura- −k33k共2兲
11 y Ⰷ 1 has to hold. In other words, the nonlinear elastoki-
2

tion, but it also remains high over a larger Y DoF displacement nematic compliance of a beam in the axial direction should be
range, as evident in Fig. 5共d兲. The use of thinner beams helps greater than its linear elastic compliance. Thus, the nondimen-
increase k33, but this by itself is not adequate because the elastoki- sional quantity −k33k共2兲 2
11 y becomes the mathematical metric for
nematic coefficient k共2兲11 also has to be large enough for elastic
determining whether elastic averaging will play a role in an over-
averaging to kick in and make the mechanism tolerant to geomet- constrained flexure mechanism design or not. This metric is listed
ric imperfections. This observation is further illustrated quantita- for several geometric configurations of the flexure mechanism un-
tively in Table 2 共Sec. 4兲, which lists several different beam ge- der consideration in Table 2, Sec. 4. This also highlights the sig-
ometries. nificance of the normalization scheme in this work, which allows
Also evident in Fig. 5共d兲 is that for a given geometric error, the a comparison of pertinent mathematical quantities and the relative
effective stiffness undergoes a greater variation in lumped- strength of physical effects.
compliance geometries. For a large enough error and degree of While k33 increases with decreasing b, k共2兲11 decreases at an even
lumped compliance 共e.g., ␣ = 0.01 and b = 0.1兲, the stiffness is pre- faster rate. However, it is the product of these two that should be

041006-6 / Vol. 2, NOVEMBER 2010 Transactions of the ASME


Capacitive
Encoder Linear
Probe (1)
Strip Mount Encoder

)
Capacitive
Elastic Averaging Metric (−k33 k(2)
11
Beams Probe (2)
Motion
Stage
Keyless
Bushing

Pulley
Beam Flexure
Construction X
Y Transmission
Line

Ground
Actuation
Frame
Weight
0
0 0.1 0.2 0.3 0.4 0.5
Fig. 7 Reconfigurable three-beam parallelogram flexure ex-
Beam Shape Parameter (b)
perimental set-up
Fig. 6 Elastic averaging metric versus beam shape

冋 共0兲
− 3k12 y+ 冉 1−
k33
共2兲 2
k33k11 y
冊 2
2 共2兲 3 2
· k11
3
y␣ 册
冋冉 冊 冉 冊 册
␪⬇ 2
high for elastic averaging to take effect. Figure 6 shows the prod- k33 k33 2 共2兲 2 2
共2兲 2 · 2w2 + 共2兲 2 · k12 y␣
uct −k33k共2兲
11 plotted against the beam shape parameter b. While the
1− k33k11 y 1− k33k11 y 3
beam thickness t affects only the vertical axis scale, which is 共16兲
intentionally omitted, the shape of this curve and location of the
Once again, the contribution of the elastokinematic nonlinearity
maxima 共b = 0.4兲 remain unchanged. This leads to the so-far un-
known and physically nonobvious conclusion that the optimal is evident here. The limiting cases of ␣ = 0, −k33k共2兲 11 y Ⰷ 1, or
2

beam shape for elastic averaging corresponds to b = 0.4 and not −k33k共2兲
11 y 2
Ⰶ 1 could be considered to further simplify the above
necessarily b = 0.5. However, since the incremental improvement expression and obtain physical insight into the behavior of this
is small, the latter may still be preferred due to ease of fabrication. parasitic rotation.
It is clear from Fig. 6 that elastic averaging will not play a sig-
nificant role for smaller values of b, which correspond to lumped-
compliance configurations. 4 Experimental Set-Up Design
5. In a design with only two beams 共exact constraint兲, where In order to experimentally validate the above analytical load-
binding is not a concern and elastic averaging is not needed, it is displacement results in a comprehensive fashion, it was deemed
noteworthy that the increase in k共0兲 11 with decreasing b is much
necessary to test several geometric configurations with varying
slower than the increase in k33, even though eventually both grow beam thicknesses 共t兲, manufacturing imperfection angles 共␣兲, and
to infinity 共see Fig. 3兲. This implies that using a beam shape degree of distributed compliance 共b兲. The measurement objective
corresponding to b = 0.25, one can increase the axial stiffness by was to record the X, Y, and ⌰ displacements of the motion stage
more than twice while increasing the transverse stiffness only by in response to known Y direction actuation forces. However, mak-
10%. However, now if one tries to increase the beam thickness to ing a separate prototype for each configuration would obviously
reduce the stresses due to axial loads or increase the axial stiff- be impractical and cost-prohibitive. Therefore, we designed a
ness, the transverse stiffness and bending stresses will increase highly reconfigurable experimental set-up, illustrated in Fig. 7,
cubically. Bending stresses, in turn, limit the transverse displace- that accommodates all of the above variations.
ment. Adding a third beam of the original thickness not only re- Blue-tempered spring steel 共ASTM A682, AISI 1095兲 shim-
stock was chosen for the beam material because of its high
duces the axial stress and increases axial stiffness, it also raises
strength to modulus ratio 共Sy = 455 MPa, E = 205 GPa, and ␯
the transverse stiffness only by 50% while keeping the bending
= 0.3兲. For a nominal beam length L = 100 mm, readily available
stresses and transverse motion range the same. Of course, to in-
thicknesses 共T = 0.2413 and 0.635 mm兲 were chosen to ensure that
corporate a third beam without causing binding, elastic averaging
a Y displacement of ⫾10 mm 共y = ⫾ 0.1兲 could be achieved with
would have to be invoked, which would require a beam with
an adequate margin of safety against yielding. For ease of fabri-
greater distributed compliance 共b ⬃ 0.4兲. In fact, more and more
cation, a standard out-of-plane dimension 共beam height兲 of H
beams can be added, while reducing the beam thickness to simul- = 25.4 mm and a beam spacing of W = 50 mm were selected.
taneously increase the axial stiffness, reduce axial stresses, keep Along with the nominal case of ␣ = 0, two geometric imperfection
the transverse stiffness and bending stresses under control, and angles, ␣ = 0.0035 and 0.007, were considered for this study. The
maximize transverse displacements. Thus, elastic averaging helps last number represents a 0.7 mm parallelism error over a 100 mm
overcome some of the fundamental performance trade-offs seen in beam length. Although most well-controlled macroscale manufac-
flexure mechanisms, and this benefit can now be quantified based turing and assembly processes are capable of producing better
on the above mathematical framework. tolerances, we decided to include potential misalignments that
For the sake of completeness, we continue from expression 共11兲 might arise in case of a human error, large volume productions, or
to find the motion stage rotation ␪, which is a parasitic error in microfabrication. Since the objective here is to understand elas-
motion in this case. Given its small magnitude, second and higher tic averaging in the context of lumped and distributed compliance,
powers of ␪ may be dropped. Setting fx = mz = 0, without loss in several levels of distributed compliance were considered 共b = 0.5,
generality, yields 0.3, 0.2, and 0.1兲. While keeping the beam spacing W, beam

Journal of Mechanisms and Robotics NOVEMBER 2010, Vol. 2 / 041006-7


Ground α - Blocks (B) Motion Stage Dowel a flat clamping surface or fixed jaw, two flexure pivots P1 and P2,
Frame Hole (A) rigid block C, rigid block D, and a pipe plug. Tightening the pipe
plug, which has tapered threads, causes block C to rotate about
pivot P1. This by itself would have provided poor clamping action
Dowel against the fixed jaw because of a localized contact. However,
Hole (B) Dowel pivot P2 in this clamp design accommodates the rotation of block
Hole (B)
C, while transmitting only a translation to block D. This allows
Flexure Clamps (B) block D, which serves as the moving jaw, to remain parallel to the
Flexure Clamp fixed jaw and provide a uniformly distributed clamping force on a
Dowel Flexure Clamps (A) P1 P2 beam that goes in between the two jaws. This results in a clamped

Fixed Jaw
Hole (A)
joint that is easy to secure or release and is free of backlash. This

Pipe Plug
C
D
clamp design also helps align the beam with respect to the fixed
jaw, while the moving jaw aligns itself to conform to the beam. A
gap of 0.75 mm was left between the fixed and moving jaws of
Beam goes each clamp to accommodate the largest beam thickness 共T
Alignment Plate α - Blocks (A) into this slot
= 0.635 mm兲.
Fig. 8 Alignment and assembly of the ground frame and mo- Next, the imperfection angle ␣ of the middle beam was
tion stage achieved using a set of four fixture blocks, referred to hereafter as
the ␣-blocks. Two ␣-blocks 共A兲 interface with the middle flexure
clamp of the ground frame, while the other two ␣-blocks 共B兲
height H, and material E invariant, the various geometric configu- interface with the middle flexure clamp of the motion stage. Each
rations that result from the above described variations are com- of these blocks have one face lined up against either the moving
piled in Table 2. or fixed jaw of its associated flexure clamp, and an opposing face
All quantities listed in this table are nondimensionalized. Values that is precision machined to be at an angle ␣ with respect to the
of the k coefficients are simply obtained from relations 共4兲 and first face. One end of the middle beam gets sandwiched between
may be seen in Fig. 3. In the final column, the product −k共2兲 2
11 k33y ,
the two angled faces of two ␣-blocks 共A兲, which in turn are sand-
which was identified as a metric for elastic averaging in Sec. 3, is wiched between the moving and fixed jaws of the middle clamp
listed for a y displacement of 0.1. The greater this number with on the ground frame. A similar arrangement is repeated on the
respect to 1, the less vulnerable is the associated geometric con- motion stage, as shown in Fig. 8. With the use of a separate set of
figuration to binding in the presence of geometric imperfections. ␣-blocks for each angle, we were able to create and test multiple
In the proposed experimental set-up, a single set of ground cases of geometric imperfection 共␣ = 0, 0.0035, and 0.007兲. For
frame and motion stage, shown in Fig. 8, accommodate all the this arrangement to work effectively, the ground frame, motion
beam variations. Three flexure clamps 共A兲 are monolithically in- stage, the ␣-blocks, and the beam flexures have to be assembled in
corporated in the ground frame and another three flexure clamps the correct position and orientation. To accomplish this, an align-
共B兲 are incorporated in the motion stage. These flexure clamps ment plate 共Fig. 8兲 was employed to build up the assembly. This
allow an easy assembly and disassembly of the individual beam plate has two pairs of dowel pins pressed in precise locations
flexures with the ground frame and motion stage. Each of these corresponding to slide-fit holes 共A兲 and 共B兲 on the ground frame
flexure clamps 共Fig. 8, inset兲, first reported in Ref. 关17兴, includes and motion stage, respectively. The first step in the assembly is to

Table 2 Geometric variations of the three-beam parallelogram flexure

␣ b k共0兲
11 k共2兲 ⴱ
11 共 10 兲
−3
k33 −k共2兲
11 k33y
2

Thickness T = 0.2413 mm, t = 0.002413


0 0.5 12.00 ⫺1.429 2,060,946 29.45
0 0.3 12.82 ⫺0.838 3,434,910 28.78
0 0.2 15.31 ⫺0.312 5,152,365 16.08
0 0.1 24.59 ⫺0.043 10,304,730 4.43
0.0035 0.5 12.00 ⫺1.429 2,060,946 29.45
0.0035 0.3 12.82 ⫺0.838 3,434,910 28.78
0.0035 0.2 15.31 ⫺0.312 5,152,365 16.08
0.0035 0.1 24.59 ⫺0.043 10,304,730 4.43
0.007 0.5 12.00 ⫺1.429 2,060,946 29.45
0.007 0.3 12.82 ⫺0.838 3,434,910 28.78
0.007 0.2 15.31 ⫺0.312 5,152,365 16.08
0.007 0.1 24.59 ⫺0.043 10,304,730 4.43

Thickness T = 0.635 mm, t = 0.00635


0 0.5 12.00 ⫺1.429 297,601 4.25
0 0.3 12.82 ⫺0.838 496,001 4.16
0 0.2 15.31 ⫺0.312 744,001 2.32
0 0.1 24.59 ⫺0.043 1,488,003 0.64
0.0035 0.5 12.00 ⫺1.429 297,601 4.25
0.0035 0.3 12.82 ⫺0.838 496,001 4.16
0.0035 0.2 15.31 ⫺0.312 744,001 2.32
0.0035 0.1 24.59 ⫺0.043 1,488,003 0.64
0.007 0.5 12.00 ⫺1.429 297,601 4.25
0.007 0.3 12.82 ⫺0.838 496,001 4.16
0.007 0.2 15.31 ⫺0.312 744,001 2.32
0.007 0.1 24.59 ⫺0.043 1,488,003 0.64

041006-8 / Vol. 2, NOVEMBER 2010 Transactions of the ASME


locate the ground frame and motion stage on the alignment plate are plotted in Fig. 9 for a representative subset of the various
via these dowel pins and holes. Next, the two outer beams are geometric configurations considered. These results reveal a good
inserted into their respective clamps, and the middle beam is in- agreement between the BCM based analytical predictions, FEA,
serted into the middle clamps along with the ␣-blocks. All these and experimental measurements. For any beam shape, for ex-
components rest on the flat alignment plate. Once the flexure ample b = 0.2 in Figs. 9共g兲, 9共d兲, and 9共a兲, it is clear that the
clamps have been tightened, the entire assembly is carefully lifted effective stiffness at small displacements 共y → 0兲 increases from a
off the alignment plate and mounted on an optics table using four nominal value with increasing ␣. However, due to elastic averag-
pillars. ing, this effective stiffness comes back down to the nominal stiff-
The ground frame, motion stage, and the ␣-blocks were made ness value, as though ␣ were zero, with increasing displacements.
from a 1 in. 共25.4 mm兲 thick AL6061-T651 plate using wire elec- This is best seen in the cases with high ␣ 共Figs. 9共a兲–9共c兲兲, where
tric discharge machining 共wire-EDM兲, to maintain dimensional, the force-displacement curve asymptotically approaches a linear
flatness, and angular tolerances associated with the flexure clamps analysis prediction at small displacements, and reverts to the
and ␣-blocks. nominal stiffness at large displacements. Both asymptotes are
Flexure beams that slide into the appropriate clamps on the highlighted via dotted lines in Figs. 9共a兲–9共c兲. A nonlinear transi-
ground frame and motion stage were designed to have a tion region between the effectiveness stiffness and nominal stiff-
Z-direction height 共H = 25.4 mm兲 that matches the height of the ness regimes is also evident.
rest of the components. To vary the beam shape, we employed For the same increase in angular imperfection, the increase
6.32 mm thick and 25.4 mm high rigid AL-6061 plates that could from nominal to effective stiffness is the more prominent for b
be bolted to the center of the spring steel strips 共Fig. 7, inset兲. For = 0.2 共Figs. 9共g兲, 9共d兲, and 9共a兲兲 compared with b = .5 共Figs. 9共i兲,
b = 0.5, the spring steel strip was used by itself for the beams, and 9共f兲, and 9共c兲兲. This confirms that lumped-compliance configura-
for all other values of b, the spring steel strip was sandwiched tions are more prone to stiffness increase, and potential binding, in
between appropriately sized rigid plates. the DoF direction. Moreover, the transition from the effective
To measure the Y displacement of the motion stage with respect stiffness regime to the nominal stiffness regime in the maximum
to the ground frame, a noncontact linear optical encoder 共 42 ␮m imperfection angle case is the smoothest for b = 0.5 共Fig. 9共c兲兲 and
resolution兲, shown in Fig. 7, was used because of its unrestricted goes through a stronger nonlinear behavior for b = .2 共Fig. 9共a兲兲.
motion range, zero-friction, simple mounting and set-up, noise- This shows that distributed compliance configurations are better at
free digital output data, and low-cost. For measuring the X and ⌰ rejecting the detrimental effects of geometric imperfection.
direction parasitic error motions of the motion stage, we employed Overall, the BCM prediction closely follows the FEA predic-
two noncontact capacitance probes mounted on the ground frame, tion for all cases, proving the effectiveness of the proposed non-
as shown in Fig. 7. These probes are ideally suited for this pur- linear closed-form parametric modeling approach. The only devia-
pose for several reasons: The range of motion to be measured is tion that occurs is in the transition region between the effective
small 共⬍0.6 mm兲, the probes are tolerant to the relatively large Y stiffness and nominal stiffness regimes for cases with high geo-
DoF motion perpendicular to their measurement axes, and the metric imperfection and low distributed compliance 共Fig. 9共a兲兲.
achievable resolution 共⬃40 nm兲 is more than adequate. Each The reason for this deviation lies in the fact that expression 共1兲 in
probe was mounted in the ground frame via a bronze bushing that the BCM is truncated to the second power of the axial load f x1 in
is split along its length and is held in place via a radial set-screw. what is originally an infinite series 关32兴. While this truncation
The bushing helps distribute the force from the set-screw uni- provides simplicity to the BCM and allows its use in the closed-
formly over the probe length, thus preventing any damage to the form parametric analysis of more complex flexure mechanisms
probe surface and associated loss of calibration. such as the one being considered here, it restricts the validity of
Y actuation of the motion stage was achieved using free hang- the model to normalized axial load values within −35⬍ f x1 ⬍ 50 to
ing weights applied in increments of 5–10 g, and measured sepa- maintain less than 2% truncation error. In fact, for larger magni-
rately to within 0.1 g. A keyless bushing was used to attach a tudes, while errors associated with tensile axial loads grow at a
nylon fishing line to one end of the motion stage; this fishing line small rate, the errors associated with compressive axial loads in-
passed through a hole on the side of the ground frame, and then crease significantly. When the three-beam parallelogram flexure is
over a low-friction pulley attached to the optics table. The keyless actuated in the positive direction 共see Fig. 4兲, it may be qualita-
bushing provides an easily attachable and detachable friction- tively argued that the middle beam will experience a compressive
based clamp that is free backlash. The motion stage and the key- axial load while the two outer beams will see the axial tensile
less bushing mounting hole in it were designed such that the loads. For the geometric configuration of Fig. 9共a兲 共large ␣ and
Y-direction actuation force would pass through the modeled point low b兲, which is most prone to overconstraint and binding, the
O on the motion stage 共see Fig. 4兲. Since free weights can provide compressive load on the middle beam grows with increasing y
actuation loads in only one direction, the same actuation set-up displacement, exceeding a normalized value of 100 in the transi-
was provided on both ends of the motion stage and ground frame tion region. This large compressive load results in the discrepancy
to allow bidirectional testing. seen between the BCM and FEA predictions. With further in-
crease in y displacement, the effective axial stiffness of the middle
beam drops because of the elastokinematic effect and therefore the
5 Analytical and Experimental Results associated compressive axial load drops. This drop ensures that
In addition to BCM based close-form analysis and experimental the BCM and FEA start to match again in the large y displacement
measurements, extensive finite element analysis 共FEA兲 was also range. When the flexure mechanism is actuated in the negative Y
conducted in ANSYS to provide yet another prediction of the three- direction, the middle beam now sees an axial tensile force while
beam parallelogram flexure’s load-displacement behavior for all the two outer beams experience axial compressive forces. How-
the geometric configurations listed in Table 2. BEAM4 elements ever, since the axial forces are equally divided between the two
were used with consistent matrix and large displacement options outer beams, their magnitude is less compared with the previous
共NLGEOM兲 turned on and shear coefficients set to zero. The case. Therefore, the agreement between the BCM and FEA is
beam flexures were meshed using up to 10 elements per mm of maintained even in the transition region in the negative Y direc-
the beam length. The convergence criterion for all cases was set to tion. This qualitatively reasoning behind the noted discrepancy
a relative tolerance limit of 0.001 on the L2 norm of the forces. has also been validated analytically.
Analytical predictions based on the BCM 共line兲 and FEA Similarly, there is an overall good agreement between the ex-
共circles兲 along with experimental measurements 共crosses兲, of the perimental measurements and BCM predictions, except for the
normalized Y actuation force versus normalized Y displacement, positive Y direction transition region. Between the experiments

Journal of Mechanisms and Robotics NOVEMBER 2010, Vol. 2 / 041006-9


5 α = 0.007 5 α = 0.007 5 α = 0.007
4 b = 0.2 4 b = 0.3 4 b = 0.5
3 3 3
2 2 2
1 1 1
0 0 0
−1 −1 −1
−2 −2 −2
−3 −3 −3
−4 −4 −4
A. B. C.
−5 −5 −5
−0.1 −0.05 0 0.05 0.1 −0.1 −0.05 0 0.05 0.1 −0.1 −0.05 0 0.05 0.1

5 α = 0.0035 5 α = 0.0035 5 α = 0.0035


4 b = 0.2 4 b = 0.3 4 b = 0.5
3 3 3
2 2 2
1 1 1
0 0 0
−1 −1 −1
−2 −2 −2
−3 −3 −3
−4 −4 −4
D. E. F.
−5 −5 −5
−0.1 −0.05 0 0.05 0.1 −0.1 −0.05 0 0.05 0.1 −0.1 −0.05 0 0.05 0.1

5 α=0 5 α=0 5 α=0


4 b = 0.2 4 b = 0.3 4 b = 0.5
3 3 3
2 2 2
1 1 1
0 0 0
−1 −1 −1
−2 −2 −2
−3 −3 −3
−4 −4 −4
G. H. I.
−5 −5 −5
−0.1 −0.05 0 0.05 0.1 −0.1 −0.05 0 0.05 0.1 −0.1 −0.05 0 0.05 0.1

Fig. 9 Normalized Y DoF force versus normalized Y DoF displacement: BCM „line…, FEA „䊊…, and Exp „ⴛ…

and FEA, the force-displacement relation trends agree for all cases prediction, FEA prediction, and experimental measurements of the
considered, with a maximum of 5% deviation in the absolute val- X and ⌰ direction parasitic error motions of the motion stage is
ues. This deviation may be attributed to an estimated value of presented in Figs. 10 and 11, respectively, for a representative
Young’s modulus used in the normalization of the experimental case 共t = 0.002413, b = 0.2, and ␣ = 0.007兲.
data-points 共E = 205 GPa兲, as opposed to an actually measured Figure 10 shows a good agreement between the BCM pre-
value. dicted, FEA predicated, and experimentally measured dependence
The FEA studies and experimental measurements serve to vali- of the X direction displacement on the Y DoF displacement. Both
date the strength and effectiveness of the BCM in capturing the the linear and quadratic components predicted by Eq. 共12兲 are
physical effects in flexures that are pertinent to the systematic and
well evident here. Deviations between analysis and experiments
quantitative study of elastic averaging. The discrepancies noted
are negligible because the relation between the plotted displace-
above do not pose a problem because they are restricted to small
regions, and the overall DoF direction force-displacement trends ments is predominantly kinematic and therefore independent of
are adequately captured. The BCM accurately predicts that even loads. Figure 11 shows a good agreement between the BCM and
for an overconstrained distributed-compliance mechanism with FEA predications of the motion stage rotation. However, the ex-
geometric imperfections, the effective DoF direction stiffness, perimental measurement of this rotation is an order of magnitude
which might be high at very small displacements, quickly reduces higher and its dependence on the Y DoF displacement is predomi-
to the nominal stiffness as though there was no geometric imper- nantly linear as opposed to the cubic prediction. This deviation is
fection. attributable to actual manufacturing tolerances in the experimental
To complete the discussion, a comparison between the BCM set-up. While the two outer beams of the flexure mechanism are

041006-10 / Vol. 2, NOVEMBER 2010 Transactions of the ASME


−3
x 10 to include overconstrained geometries that might be capable of
1 higher performance but are traditionally ruled out by the exact
α = 0.007 constraint design principles.
Normalized Axial Displacement (x) 0 b = 0.2 Quantitatively, the key highlights of this paper are as follows:
共1兲 A linear analysis shows that geometric imperfections affect
−1 lumped-compliance configurations more in terms of an increased
effective stiffness in the DoF direction, resulting in a potential
−2 mobility loss. 共2兲 Nonlinear analysis is needed to show that the
increased effective stiffness comes back to the nominal stiffness
−3 with increasing displacements, as though there were no geometric
imperfections. This transition is faster and smoother for distrib-
−4
uted compliance configurations. 共3兲 The nonlinear elastokinematic
effect, along with the linear axial stiffness of the flexure beam,
helps define a metric for elastic averaging. The larger this nondi-
−5
mensional number compared with 1, the greater is the resulting
mechanism’s ability to reject the detrimental effects of geometric
−6 imperfections. 共4兲 The nonlinear closed-form load-displacement
model of the three-beam parallelogram flexure mechanism is
−7 based on the simple yet powerful BCM, which captures elastic,
−0.1 −0.05 0 0.05 0.1
Normalized DoF Displacement (y ) load-stiffening, kinematic, and elastokinematic effects. A general-
ized BCM that accommodates any beam shape is used here to
Fig. 10 Normalized X displacement: BCM „line…, FEA „䊊…, and map the degree of distributed compliance to elastic averaging ca-
Exp „ⴛ… pability. 共5兲 An energy approach is presented that simplifies the
mathematical derivation of force-displacement relations for the
mechanism from the BCM. Although a three-beam parallelogram
assumed to be perfectly parallel to the X axis and equidistant from mechanism and a simple parallelism error were considered here to
point O, some deviation, however, small, will still exist in spite of highlight the concept of elastic averaging, the present approach
wire-EDM fabrication and careful assembly. It may be analyti- can be readily extended to a parallelogram mechanism with any
cally shown that while such small deviations do not affect the Y number of beams and other types of geometric errors. 共6兲 The
and X direction results, they can significantly affect the ⌰ rota- BCM based closed-form predictions are validated via experimen-
tion, which is a much smaller quantity 共⬃10−4兲 and far more tal measurements and FEA.
sensitive to the mechanism geometry. An unaccounted angular
deviation of 0.0005 rad from perfect zero 共which is in the range of Acknowledgment
wire-EDM tolerances兲 in either of the outer two beams can lead to This research was supported in part by a National Science
a dominant linear effect in the motion stage rotation, of the order Foundation grant 共Grant No. CMMI 0846738兲.
of what is seen in the experimental measurements 关18兴.

6 Conclusion References
关1兴 Maxwell, J. C., 1980, The Scientific Papers of James Clerk Maxwell, Vol. 2,
This paper provides a systematic mathematical basis for study- Cambridge University Press, London.
ing and employing elastic averaging in flexure mechanism design. 关2兴 Pollard, A. F. C., 1929, The Kinematic Design of Couplings in Instrument
As a design paradigm, elastic averaging is particularly suited to Design, Hilger and Watts, London.
distributed compliance flexure mechanisms because of their finite 关3兴 Whitehead, T. N., 1934, The Design and Use of Instruments and Accurate
Mechanisms, Macmillan, New York.
constraint direction stiffness, which makes them tolerant to typical 关4兴 Furse, J. E., 1981, “Kinematic Design of Fine Mechanisms in Instruments,” J.
manufacturing and assembly errors. This expands the design space Phys. E, 14, pp. 264–272.
关5兴 Erdman, A. G., Sandor, G. N., and Kota, S., 2001, Mechanism Design: Analy-
sis and Synthesis, Prentice-Hall, Englewood Cliffs, NJ.
−4 关6兴 Ullman, D. G., 2003, The Mechanical Design Process, 3rd ed., McGraw-Hill,
x 10 New York.
4
关7兴 Whitney, D. E., 2004, Mechanical Assemblies: Their Design, Manufacture,
α = 0.007 and Role in Product Development, Oxford University Press, New York.
3 b = 0.2 关8兴 Whitehouse, D. J., 2003, Handbook of Surface and Nanometrology, Institute of
Physics, Bristol.
关9兴 Moore, W. R., 1970, Foundations of Mechanical Accuracy, Moore Tool Co.,
2
Motion Stage Rotation (θ )

Bridgeport, CT.
关10兴 Skakoon, J. G., 2008, The Elements of Mechanical Design, ASME, New York.
1 关11兴 Slocum, A. H., 1992, Precision Machine Design, Society of Manufacturing
Engineers, Dearborn, MI.
关12兴 Hale, L. C., 1999, “Principles and Techniques for Designing Precision Ma-
0 chines,” Ph.D. thesis, Massachusetts Institute of Technology, Cambridge, MA.
关13兴 Blanding, D. K., 1999, Exact Constraint: Machine Design Using Kinematic
Principles, ASME, New York.
−1 关14兴 Grubler, M., 1917, Getriebelehre, Springer, Berlin, Germany.
关15兴 Hopkins, J. B., 2005, “Design of Parallel Systems via Freedom and Constraint
−2 Topologies 共FACT兲,” MS thesis, Massachusetts Institute of Technology, Cam-
bridge, MA.
关16兴 Jones, R. V., 1988, Instruments and Experiences: Papers on Measurement and
−3 Instrument Design, Wiley, New York.
关17兴 Awtar, S., 2004, Analysis and Synthesis of Planer Kinematic XY Mechanisms,
Sc.D. thesis, Massachusetts Institute of Technology, Cambridge, MA.
−4 关18兴 Awtar, S., Slocum, A. H., and Sevincer, E., 2007, “Characteristics of Beam-
−0.1 −0.05 0 0.05 0.1
Normalized DoF Displacement (y ) Based Flexure Modules,” ASME J. Mech. Des., 129共6兲, pp. 625–639.
关19兴 Bonin, W. A., 2001, “Microactuator Suspension With Multiple Narrow
Beams,” U.S. Patent No. 6,282,066.
Fig. 11 Motion stage rotation: BCM „line…, FEA „䊊…, and Exp 关20兴 Trease, B. P., Moon, Y.-M., and Kota, S., 2005, “Design of Large Displace-
„ⴛ… ment Compliant Joints,” ASME J. Mech. Des., 127, pp. 788–798.

Journal of Mechanisms and Robotics NOVEMBER 2010, Vol. 2 / 041006-11


关21兴 Awtar, S., and Sevincer, E., 2006, “Elastic Averaging in Flexure Mechanisms: 关28兴 Jiang, L., 2007, “A Novel Method for Nanoprecision Alignment in Wafer
A Multi-Parallelogram Flexure Case-Study,” ASME Paper No. 99752. Bonding Application,” J. Micromech. Microeng., 17共7兲, pp. S61–S67.
关22兴 Wilson, E. B., Jr., 1952, An Introduction to Scientific Research, McGraw-Hill, 关29兴 Jurcevich, B. K., 1992, “Structural Mechanics of the Solar—A Soft X-Ray
New York. Telescope,” Proc. SPIE, 1690, pp. 399–434.
关23兴 Jones, R. V., 1962, “Some Uses of Elasticity in Instrument Design,” J. Sci. 关30兴 Awtar, S., and Slocum, A. H., 2007, “Constraint-Based Design of Parallel
Instrum., 39, pp. 193–203. Kinematic XY Flexure Mechanisms,” ASME J. Mech. Des., 129共8兲, pp. 816–
关24兴 Gardner, J. W., and Hingle, H. T., 1991, From Instrumentation to Nanotech- 830.
nology, Gordon Breach, Philadelphia. 关31兴 Awtar, S., and Sen, S., 2010, “A Generalized Constraint Model for Two-
关25兴 Willoughby, P., 2005, “Elastically Averaged Precision Alignment,” Ph.D. the- Dimensional Beam Flexures: Non-Linear Load-Displacement Formulation,”
sis, Massachusetts Institute of Technology, Cambridge, MA. ASME J. Mech. Des., 132共8兲, p. 081008.
关26兴 Slocum, A. H., and Weber, A. C., 2003, “Precision Passive Mechanical Align- 关32兴 Awtar, S., and Sen, S., 2010, “A Generalized Constraint Model for Two-
ment of Wafers,” J. Microelectromech. Syst., 12共6兲, pp. 826–834. dimensional Beam Flexures: Nonlinear Strain Energy Formulation,” ASME J.
关27兴 Slocum, A. H., Basaran, M., Cortesi, R., and Anastasios, J. H., 2003, “Linear Mech. Des., 132共8兲, p. 081009.
Motion Carriage With Aerostatic Bearings Preloaded by Inclined Iron Core 关33兴 Oh, Y. S., 2008, “Synthesis of Multistable Equilibrium Compliant Mecha-
Linear Electric Motor,” Precis. Eng., 27, pp. 382–394. nisms,” Ph.D. thesis, University of Michigan, Ann Arbor, MI.

041006-12 / Vol. 2, NOVEMBER 2010 Transactions of the ASME

You might also like