You are on page 1of 22

Mechanism

and
Machine Theory
Mechanism and Machine Theory 39 (2004) 921–942
www.elsevier.com/locate/mechmt

A comparison of the performances of full and half


toroidal traction drives
G. Carbone *, L. Mangialardi, G. Mantriota
Dipartimento di Ingegneria Meccanica e Gestionale, Politecnico di Bari, V.le Japigia 182, 70126 Bari, Italy
Received 5 May 2003; received in revised form 21 January 2004; accepted 10 April 2004

Abstract
The efficiency of two different typologies of the toroidal traction drive, the full-toroidal and the half-
toroidal, is estimated in order to point out which of them offers the higher mechanical efficiency. A fully
flooded isothermal contact model between the discs and rollers, based on the results of EHL theory, is used
to evaluate the slip, the spin losses and the mechanical performances of the variators. It is shown that the
half-toroidal traction drive offers higher efficiency and higher maximum transmissible torque.
Ó 2004 Elsevier Ltd. All rights reserved.

1. Introduction

The request for a higher energy efficiency and CO2 reduction has pushed several researchers to
find new technical solutions to improve the emission performance of nowadays IC engine vehicles.
While waiting for new and renewable forms of energy to become effective and cost reasonable,
new solutions have to be found: among different and several technical solutions, new drive train
systems are being investigated to accomplish this purpose. The continuously variable transmission
(CVT) represents one of the most promising solution since it is able to provide an infinite number
of gear ratios between two finite limits, and, thus, to allow the IC engine to operate closer to its
optimal efficiency line.
Several studies have shown that it is possible to improve the fuel savings and to reduce the
vehicle emissions by adopting this kind of transmission [1,2]. It has been shown, for instance, that
when the transmission is optimally controlled the mid class vehicles equipped with the CVT and

*
Corresponding author. Tel.: +39-080-596-2746; fax: +39-080-596-2777.
E-mail addresses: carbone@poliba.it (G. Carbone), lmm@poliba.it (L. Mangialardi), mantriota@poliba.it (G.
Mantriota).

0094-114X/$ - see front matter Ó 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mechmachtheory.2004.04.003
922 G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942

IVT variators may achieve less fuel consumption (about 10% for CVT and 6% for IVT) and
higher comfort in the urban traffic [3]. Moreover applications of CVTs to wind power systems
have been proposed, and some papers [4–6] have shown that a significant increase of the energy
production may be attained. The metal pushing V-belt and the metal chain CVTs are able to
achieve these results, but they have some drawbacks as the strict dependence of the shifting speed
on the clamping forces acting on the moving pulley sheaves [7–9], and the smaller torque capacity
when compared to toroidal traction drives.
The robotized gearbox may be a different solution, since it combines the economy of a well-
driven manual transmission with the easy to use of a conventional automatic transmission, and
get the further advantage to retain the simplicity and the economies of scale of an established
manual design. But, in this case, some problems regarding the shift quality arise: the torque
interruption becomes intrusive because the driver is not able to predict or anticipate the gear shift.
For these reasons more promising solutions have to be found, and the toroidal traction drives
may be one of these. They are being extensively investigated because of their high torque capacity,
that makes them suitable for application in larger engine cars and even trucks. The most attractive
typologies are the full-toroidal [10,11] and the half-toroidal traction drives [12–14]. The main
components of these transmissions are the input and output discs, designed to create a toroidal
cavity (see Fig. 1), coupled with an appropriate number of rollers. The high torque capacity of
these transmissions is obtained by coupling together in a series scheme two or more single units
[15–17]. Moreover the particular geometry of the toroidal traction drive makes it able to rapidly
adjust its speed ratio to the request of the driver, thus improving the driving comfort [18–21].
Between the roller and the discs no metal–metal contact occurs, the torque is transmitted by
means of the shearing action of a special oil referred to as traction oil. The lubrication regime of
such a system is the hard EHL with pressures up to 3 GPa. The high pressures lead to a much
higher oil viscosity (several order of magnitude) than in the normal hydrodynamic regime, thus
enabling the transmission of high torque despite the very small area of contact.
In the technical literature many works can be found concerned with the design, fatigue life, and
thermal stresses of the half-toroidal CVT [22–26], but not many contributions compare the full-

(a) (b)

Fig. 1. The traction drive CVTs: (a) half-toroidal; (b) full toroidal.
G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942 923

Table 1
CVT geometric data
Half toroidal CVT: Full toroidal CVT
Cavity radius r0 ¼ 40 mm Cavity radius r0 ¼ 40 mm
Roller curvature r22 ¼ 32 mm Roller curvature r22 ¼ 26:4 mm
Half cone-angle h ¼ p=3 Half cone-angle h ¼ p=2
Aspect ratio k ¼ e=r0 ¼ 0:625 Aspect ratio k ¼ e=r0 ¼ 0:25
Speed ratio range s ¼ 0:5–2:0 Speed ratio range s ¼ 0:5–2:0

toroidal and half-toroidal traction drives as regards the mechanical efficiency and their traction
capabilities. Only few papers provide some experimental results on their efficiency [27–29].
The main scope of this work is to propose a theoretical model of the variators, based on the
results of EHL lubrication theory, that is able to estimate the traction capability and the
mechanical efficiency of these two different CVT typologies, and to point out which of them offers
the higher mechanical efficiency. The analysis is limited to the simple variator as it is, made up of
the input discs, the rollers, the support bearings and the output discs. The two variator are
supposed to have the same cavity radius and the same radial size, the latter condition requires
different values of the aspect ratio k ¼ e=r0 as reported in Table 1. The principal advantage of the
proposed method with respect to other similar methodologies [30,31] consists of three points: the
model is independent of the specific traction drive under consideration, the formulation presented
does not require the determination of the traction curve by experiments performed on the given
traction drive with the specific traction oil. The model is also able to take into account the
influence of the spin motion on the mechanical efficiency and the traction performance of the
variator. The main drawback is related to the large number of equations to deal with, that makes
the overall computation time-consuming.

2. Geometric and kinematic quantities of full and half toroidal CVTs

2.1. Geometrical description of CVTs

Fig. 1 shows the main geometrical features of the toroidal variators. During the steady state
operation of the CVT the swing center of the roller coincides with the cavity center O, and its axis
of rotation is tilted of c. The tilting angle c (positive if clockwise directed) controls the distance r1
and r3 of the contact points A and B from the main axis of the variator, and, consequently,
controls the ideal speed ratio srID ¼ r3 =r1 . In the same Fig. 1, r12 ¼ r23 ¼ r0 represent radius of the
toroidal cavity, that is also one of the two principal radii of curvature of the input and output
discs. The quantity r11 is the second principal radius of curvature of the input disc, whereas r33 is
the second principal radius of curvature of the output disc. Moreover r2 and r22 are the two
principal radii of curvature of the roller with r22 < r12 . The quantity e is the distance of the
toroidal cavity from the disc axes, it is related to the aspect ratio k ¼ e=r0 of the toroidal traction
drive. Moreover in a half-toroidal CVT, the half cone-angle h of the roller is about 50–70°,
whereas in a full-toroidal CVT the cone angle is 90°. All the remaining geometrical parameters are
reported in Table 1.
924 G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942

2.2. Kinematic analysis of CVTs

In this subsection the kinematics of the toroidal traction drives is analyzed during the steady
state operation of the variator. Consider the roller and the discs as rigid bodies and assume no-slip
at points A and B (see Fig. 2). Under these conditions, the motion of the roller relative to the input
disc is a spherical rigid motion, of which the instantaneous axis of rotation can be easily deter-
mined as the straight-line through the points of null relative velocity A and X (the point X is the
intersection of the roller absolute axis of rotation and the input-disc absolute axis of rotation).
Similar arguments hold when studying the relative velocity field between the roller and the output-
disc, in this case the instantaneous axis of rotation of the relative motion is the straight-line BX.
Let x1 , x2 and x3 be, respectively, the absolute angular velocities of the input disc, of the roller
and that of the output disc. The angular velocity of the roller relative to the input disc is,
therefore, x21 ¼ x2  x1 , whereas that one relative to the output disc is x23 ¼ x2  x3 . Figs. 2
and 3 show these two relative velocities of rotation x21 and x23 both for the half-toroidal and full-
toroidal CVTs. It is clearly shown that, since the point of intersection H (see Fig. 2) of the two
tangents to the toroidal cavity at points A and B does not always coincides with the point X, both
the relative angular velocities x21 and x23 have non-zero spin vector components x21spin and
x23spin , respectively. Moreover, the ratio ðx21spin Þin =jx21 j ¼ ðx23spin Þout =jx23 j ¼ sin a, assumes its
maximum value for srID ¼ r3 =r1 ¼ 1, since the distance between the points H and X is maximum
for this value of the ideal speed ratio. Furthermore, Fig. 2 shows that, for the half-toroidal CVT,
two points exist at which H and X coincide and the spin motion vanishes.
Different considerations have to be done for the full-toroidal traction drive. This time, Fig. 3
shows that the point H goes to infinity, thus the spin motion never vanishes, and, because of the
bigger angle a, it is always bigger than in the case of the half-toroidal CVT. Once again, the worst
situation occurs for srID ¼ 1, when (see Fig. 3) the modulus of the spin vector components x21spin
and x23spin equals the modulus of absolute angular velocity of the output and input discs,
respectively.
What happens when slip occurs, namely when torque is transmitted, is only slightly different
from the above written scenario since the slip is always very small and, hence, the axes of rotation

(a) (b)

Fig. 2. The half toroidal CVT: spin motion and rolling motion of the roller depicted for no-slip conditions; (a) speed
ratio equal to 1; (b) no spin condition (speed ratio differs from the unit value).
G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942 925

(a) (b)

Fig. 3. The full-toroidal traction drive: (a) speed ratio equal to 1; (b) same speed ratio as in (Fig. 2(b)) but the spin is
always different from zero.

of the roller relative to the discs would result only slightly tilted relatively to those depicted in the
Figs. 2 and 3. Now consider the contact area between the roller and the discs. In this region the
elastic deformations of the bodies have a large influence on the relative velocity motion, and this
one can no more be classified as a rigid motion. The region of contact is an elliptical area centered
at the point of contact. The ellipse principal axes, (see Fig. 4) lay on the y-axis (the major) and on
the x-axis (rolling direction, the minor). In order to calculate the shear strain of the lubricant we
need to find an explicit formulation of the relative velocity field in the contact region. Observe that,
over the contact area the bodies cannot penetrate each other, thus the relative velocity, because of
the elastic deformations, do not have any component normal to the area of contact. Hence,
assuming a negligible tangential deformation of the elastic bodies, the velocity of the roller points
relative to the input and output discs, respectively, can be written, in the region of contact, as:
v21 ¼ v21A þ x21spin ^ ðPin  AÞ ð1Þ
v23 ¼ v23B þ x23spin ^ ðPout  BÞ ð2Þ
Pin and Pout are points of the roller, while the velocity vectors v21A and v23B stand for the relative
velocity between roller and discs at the center points A and B of the contact areas. Since we are
considering only steady-state behavior v21A and v23B do not have components along the y-axes, yin
and yout (see Fig. 4), but only along the x-axes. The previous Eqs. (1) and (2) show that the relative
motion between the roller and the discs in the contact region, can be split into a pure translation,
given by the vectors v21A and v23B , and a pure spin motion about the z-axes.

2.3. Practical aspects for CVTs

When studying the toroidal traction drives, it is useful to define the input and output slip
coefficients, usually referred to as creep coefficients Crin and Crout :
jx1 jr1  jx2 jr2 jx2 jr2  jx3 jr3
Crin ¼ ; Crout ¼ ð3Þ
jx1 jr1 jx2 jr2
926 G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942

Fig. 4. The reference frames used at the input and output points of contact.

A small amount of creep must be always present to allow the transmission of torque. Besides the
creep coefficients defined above, it is useful, for the next calculations, to introduce the following
dimensionless geometric quantities (remember that r12 ¼ r23 ¼ r0 , see also Fig. 1):
r1
~r1 ¼ ¼ 1 þ k  cosðh þ cÞ
r0
r3 ð4Þ
~r3 ¼ ¼ 1 þ k  cosðh  cÞ
r0
By means of the creep coefficients and considering that srID ¼ ~r3 =~r1 it is possible to write the actual
speed ratio sr ¼ jx3 j=jx1 j as:
1 þ k  cosðh þ cÞ
sr ¼ ð1  Crin Þð1  Crout Þ ¼ ð1  Crin Þð1  Crout ÞsrID ð5Þ
1 þ k  cosðh  cÞ
and also define the speed efficiency mspeed of the variator as the ratio sr =srID :
sr
mspeed ¼ ¼ 1  Cr ð6Þ
srID
where 1  Cr ¼ ð1  Crin Þð1  Crout Þ stands for the global sliding coefficient between the output
disc and input one. In a similar way, it is also possible to write the spin ratios as a function of the
creep coefficients:
1 þ k  cosðh þ cÞ
r21 ¼ sinðh þ cÞ  ð1  Crin Þ ð7Þ
tan h
1 1 þ k  cosðh  cÞ
r23 ¼ sinðh  cÞ  ð8Þ
1  Crout tan h
G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942 927

1.1
(σ 21)no-slip
1
0.9
0.8
0.7
0.6
0.5 Full Toroidal
Half Toroidal
0.4
0.3

0.2
0.1
0

0.5 0.75 1 1.25 1.5 1.75 s r ID 2

Fig. 5. The spin-ratios as a function of the ideal speed ratio srID , for no-slip conditions (traction drive data in Table 1).

The above written Eqs. (4)–(8) hold true also for the full toroidal traction drive with h ¼ p=2. As
already discussed before, it is shown that, for the ideal case of no-slip, the worst condition as
regards the magnitude of the spin occurs for c ¼ 0, that is to say for srID ¼ 1. In fact, replacing
both Crin and Crout by zero, Eqs. (7) and (8) become:
cos c  ð1 þ kÞ cos h
ðr21 Þno-slip ¼ ¼ ðr23 Þno-slip ð9Þ
sin h
Equation (9) shows that the two spin ratios r21 and r23 assume their maximum value when
cos c ¼ 1, i.e. c ¼ 0. Observe that if cos h < ð1 þ kÞ1 two different values of the tilting angle c also
exist at which the spin ratios vanish, as pointed out in Section 2.2 (see also Fig. 2).
For full toroidal CVT replacing h by p=2 in Eq. (9) we obtain ðr21FT Þno-slip ¼ ðr23FT Þno-slip ¼ cos c
which is always bigger than the spin ratios of an half-toroidal CVT. Fig. 5 shows, for no-slip
conditions, the spin-ratios as a function of the ideal speed ratio srID . The CVT geometrical
characteristics are reported in Table 1, where the radii of curvature r22 have been chosen in order
to obtain the same maximum shear stress in both CVTs (see also Section 4.1). As before predicted
the spin ratio of the full-toroidal CVT is about five times higher than that of the half-toroidal one.

3. Forces, spin momentum and efficiency

The presence of spin motion affects the full-toroidal traction drive more than the half-toroidal
variator. But, on the other hand, the latter is affected by the support bearing losses, since the
normal forces FN acting on the roller, at the points of contact, do not balance out (see Fig. 6).
Therefore, a resulting axial force FR has to be supported by an axial bearing (one for each roller),
that, because of its internal losses, causes a reduction of the CVT mechanical efficiency. On the
other hand, the full-toroidal variator is not affected by this problem, since h ¼ p=2, and the
normal forces balance out. But, on the other hand, the full-toroidal CVT is much more affected by
spin losses, thus it is necessary to carry out a comparison in order to single out which typology of
CVT offers the higher mechanical efficiency.
928 G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942

Fig. 6. The free body diagram of the discs and roller; FN is the normal force at the point of contact, FR is the resulting
load on the axial bearing, FDin and FDout are the axial clamping forces on the input and output discs, respectively, FTin and
FTout are the traction (tangential) at the input and output points of contact, respectively, Tin and Tout are the input and
output torques, MSin and MSout are the spin momenta at the input and output points of contact, respectively, TBL is the
torque resistance of the axial bearing, n is the number of rollers per each cavity, and m is the number of cavities.

As regards the support bearing losses some studies have been carried out on this aspect of the
HT-traction drives. For example in [25] the authors estimate the spinning losses of the power
roller bearing by means of an elastic–plastic lubricant model, and obtain results in good agree-
ment with the experiments. For our scope we will make use of the empirical relation Eq. (10),
already used in [17], that gives the torque loss as a function of the axial thrust acting on the roller
FR , and that results in agreement with the experimental results reported in [25]. In Eq. (10) the
axial thrust acting on the roller FR is measured in (N) and the torque bearing loss TBL is measured
in (Nm).
TBL ¼ 4:6  105 FR1:03 ð10Þ
Fig. 6 shows the free body diagram of the input and output discs, and that one of the roller. The
support bearing is modelled as a revolute movable joint, that prevents the translatory motion of
the roller along the direction of its axis of rotation. The force balance of the roller gives:
FR ¼ 2FN cos h ð11Þ
and
FTin r2  FTout r2  TBL þ MSin cos h þ MSout cos h ¼ 0 ð12Þ
where FTin and FTout stand for the traction forces at the points of contact, and MSin and MSout are the
corresponding spin momenta.
The equilibrium of the input and output discs, once chosen the number of rollers n and the
number of cavity m, gives:
FDin ¼ nFN sinðh þ cÞ ð13Þ
G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942 929

Tin =m ¼ nFTin r1 þ nðMspin Þin sinðh þ cÞ ð14Þ


FDout ¼ nFN sinðh  cÞ ð15Þ
Tout =m ¼ nFTout r3  nðMspin Þout sinðh  cÞ ð16Þ
where Tin and Tout stand for the input and output torque, and FDin and FDout stand for the axial load
on the input and output discs.
Let us now define the traction coefficient l as the ratio between the traction tangential force FT
and the normal force FN , at the input and output sides of the variator:
lin ¼ FTin =FN
ð17Þ
lout ¼ FTout =FN
Also the spin momentum coefficients vin and vout can be defined as:
vin ¼ MSin =ðFN r1 Þ
ð18Þ
vout ¼ MSout =ðFN r3 Þ
Together with these coefficient it is useful to define the following dimensionless quantities:
fDin ¼ FDin =ðnFN Þ; fDout ¼ FDout =ðnFN Þ
fR ¼ FR =FN
ð19Þ
tin ¼ Tin =ðmnFN r1 Þ; tout ¼ Tout =ðmnFN r3 Þ
tBL ¼ TBL =ðFN r0 Þ
With the above mentioned dimensionless quantities, Eqs. (11)–(16) can be rephrased in a
dimensionless form as:
fR ¼ 2 cos h ð20Þ
tBL ¼ ðlin  lout Þ sin h þ fvin ½1 þ k  cosðh þ cÞ þ vout ½1 þ k  cosðh  cÞg cos h ð21Þ
fDin ¼ sinðh þ cÞ; fDout ¼ sinðh  cÞ ð22Þ
tin ¼ lin þ vin sinðh þ cÞ; tout ¼ lout  vout sinðh  cÞ ð23Þ
The above defined quantities, enable us to find a simple expression of the mechanical efficiency of
the variators:
Tout x3 r1 Tout r3 x3 r1 Tout sr r1 Tout
m¼ ¼ ¼ ¼ mspeed ¼ mtorque mspeed
Tin x1 r3 Tin r1 x1 r3 Tin srID r3 Tin
where the torque efficiency mtorque has been defined as:
ðTout =r3 Þ tout lout  vout sinðh  cÞ
mtorque ¼ ¼ ¼ ð24Þ
ðTin =r1 Þ tin lin þ vin sinðh þ cÞ
It represents the ratio between the actual output torque Tout and the output torque that would be
transmitted if the the spin momenta MSin and MSout and the torque loss in the roller bearing TBL
were absent. The overall CVT mechanical efficiency can, therefore, be rewritten as:
930 G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942

lout  vout sinðh  cÞ


m ¼ mspeed mtorque ¼ ð1  CrÞ ð25Þ
lin þ vin sinðh þ cÞ

4. Contact model

A fully flooded isothermal contact model between disks and rollers, based on the results of
EHL theory, is adopted to evaluate the slip and spin losses. Because of the severe fluid contact
conditions the Bair and Winer non-Newtonian model is used to describe the rheological behavior
of the fluid. The influence of the pressure on the limiting shear stress is also taken into account,
and according to Roelands [32], the effect of the pressure on the fluid viscosity is considered too.
The film thickness of the traction oil is estimated by means of the Hamrock and Dowson formulas
[32,33], whereas the pressure distribution over the contact area is supposed to obey to the Hertz
law for the dry contact. This last hypothesis is commonly adopted in hard-EHL contacts because
the very high contact pressure results in an almost constant thickness of the oil film, except for a
very narrow area near to the outlet region of the contact [34–36].
The model does not account for the influence of the temperature gradients on the fluid prop-
erties, since there is not an universally accepted technique to calculate this effect. The preferred
methods are based on the fluid flash temperature, but these techniques have been developed for
line contacts, and they are very difficult to validate experimentally since there is no direct way to
measure the fluid temperature. Therefore, it is expected that, for high values of the creep coeffi-
cients, the performances of the traction drives will be worse than those calculated by the proposed
model. But, typically, the creep coefficient are limited to 2–3% and the proposed analysis can be
still considered accurate.

4.1. Contact pressure distribution

The evaluation of the pressure distribution and the calculation of the extension of the contact
region need the knowledge of the equivalent radius of curvature of the contacting surfaces.
Moreover, also the ellipticity parameter e and the complete elliptic integrals of the first and second
kinds I1 and I2 need to be known. The equivalent curvature of the contacting surfaces is easy to
calculate since the geometry of the variators is given. Therefore, once defined the dimensionless
radii of curvature as the ratio of the dimensional radius of curvature and the cavity radius r0 , i.e.
~j ¼ qj =r0 and ~r22 ¼ r22 =r0 (the subscript j refers to the generic radius of curvature) the following
q
relations hold true at the input and output points of contact, respectively:

1 r0 1þk 1 r0 1þk
¼ ¼ ; ¼ ¼ ð26Þ
q
~eqX qeqX 1 þ k  cosðh þ cÞ q
~eqX qeqX 1 þ k  cosðh  cÞ
in in out out

1 r0 1 1 r0 1
¼ ¼  1; ¼ ¼ 1 ð27Þ
q
~eqY qeqY ~r22 q
~eqY qeqY ~r22
in in out out

1 r0 cosðh þ cÞ 1 1 r0 cosðh  cÞ 1
¼ ¼ þ ; ¼ ¼ þ ð28Þ
~eqin
q qeqin 1 þ k  cosðh þ cÞ ~r22 ~eqout
q qeqout 1 þ k  cosðh  cÞ ~r22
G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942 931

To evaluate the pressure distribution, the semi-axes aX and aY of the contact ellipse have to be
calculated, the simplified approach of Hamrock and Brewe [32] is adopted, thus the ellipticity
R p=2  1=2
parameter e ¼ aY =aX , and the elliptic integrals I1 ¼ 0 1  ð1  1=e2 Þ sin2 / d/ and
R p=2  2 2
1=2
I2 ¼ 0 1  ð1  1=e Þ sin / d/ can be estimated as [32]:
e ¼ n2=p ð29Þ
p p 
I1 ¼ þ  1 ln n ð30Þ
2 2
p 1
I2 ¼ 1 þ 1 ð31Þ
2 n
where the dimensionless quantity n stands for the ratio between the principal radii of curvature
n¼q qeqX .
~eqY =~
The calculation of the semi-axes aX and aY of the contact ellipse can be done, now, by means of
Eq. (34) where the dimensionless semi-axes ~aX and ~aY , defined in Eq. (33) appear.
The contact length parameter is:
 1=3
6FN r0
K¼ ð32Þ
pE0
and the dimensionless semi-axis of the contact ellipse are:
aX aY
aX ¼ ; ~
~ aY ¼ ð33Þ
K K
where the quantity E0 is the effective elastic modulus defined as E0 ¼ E=ð1  m2 Þ, m is the Poisson’s
ratio and E is the modulus of elasticity of both roller and disc. The Hertz formulas give [32]:
!1=3
 1=3 I q
~
2 eq
aY ¼ e2 I2 q
~ ~eq ; ~aX ¼ ð34Þ
e

By introducing the dimensionless pressure p~ ¼ pK2 =FN , the dimensionless maximum half-ampli-
tude of the subsurface orthogonal shear stress ~s0 ¼ s0 K2 =FN and the dimensionless co-ordinates
X ¼ x=aX , Y ¼ y=aY the Hertz theory yields:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3 1
p¼~
~ pMax 1  X 2  Y 2 ; ~ pMax ¼ ð35Þ
2 p~
aX ~
aY
ð21  1Þ1=2
~s0 ¼ ~
pMax ð36Þ
21ð1 þ 1Þ
where the auxiliary quantity 1 satisfies the following relation:
e2 ð12  1Þð21  1Þ  1 ¼ 0 ð37Þ
Fig. 7 shows that, by means of an appropriate choice of the rollers curvature r22 , it is possible to
obtain, over the whole ratio range, almost the same value of ~s0 ¼ s0 K2 =FN for both variators
investigated in this paper. Observe that this quantity is one of the most important parameter to
evaluate the stress severity as regards the fatigue life of the variator. A good choice consists in
932 G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942

(τ 0 Λ2)/F N
0.4

0.36

0.32

0.28
Half-Toroidal
0.24 Full-Toroidal

0.2
0.5 0.75 1 1.25 1.5 1.75 2
s r ID

Fig. 7. The dimensionless maximum half-amplitude of the subsurface orthogonal shear stress ~s0 ¼ s0 K2 =FN versus the
ideal speed ratio srID for the following dimensionless parameters ~r22HT ¼ r22HT =r0 ¼ 0:8, ~r22FT ¼ r22FT =r0 ¼ 0:66.

making ~s0 exactly the same when the ideal speed ratio is srID ¼ 1. To obtain this result the fol-
lowing values of the dimensionless parameter ~r22 ¼ r22 =r0 has been chosen:
~r22HT ¼ r22HT =r0 ¼ 0:8; ~r22FT ¼ r22FT =r0 ¼ 0:66 ð38Þ

4.2. EH lubrication model

In EHL contacts the pressure can rise up to 3 GPa, thus producing very severe lubricant
operative conditions. According to the Roelands model [32] and considering the isothermal
contact hypothesis the following Eq. (39) enables us to estimate the fluid viscosity over the whole
contact region:
  " !Z 1 #  
g p~
p g0
gÞ ¼ log
logð~ ¼ 1þ  1 log ð39Þ
g0 6R~cp g1

where ~cp ¼ cp =E0 , cp ¼ 1:96  108 Pa, g is the absolutely viscosity at the pressure p, g0 is the
absolute viscosity at the atmospheric pressure for the given temperature, g1 ¼ 6:31  105 Pa s,
the dimensionless constant Z1 is the viscosity-pressure index and the new dimensionless load
parameter is:
 0 2 1=3
r0 pE r0
R¼ ¼ ð40Þ
K 6FN
The non linear behaviour of the traction oil is described in Eq. (41) according to the normal rule
used in the plasticity theory to split the shear strain along the different directions:
ovi ovj sij
þ ¼ Cðse Þ ð41Þ
oxj oxi se
G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942 933

In the above written Eq. (41) the equivalent stress se has been defined as:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
se ¼ ðsij  sij Þ=2 ð42Þ
The function Cðse Þ is representative of the non-linear behavior of the oil. One of the most
commonly used explicit representation of this function is the one by Bair and Winer [32]:
 
sL 1
Cðse Þ ¼ ln ð43Þ
g 1  se =sL
The quantity sL is the limiting shear stress of the lubricant and is normally evaluated, for a certain
value of the oil temperature, by means of the following Eq. (44):
sL ¼ sL0 þ ap ð44Þ
where sL0 is the limiting shear stress at the atmospheric pressure [32].
Observe that, on that part of the contact region where significant shear stresses are involved, the
film thickness of the lubricant could be considered almost constant, thus the following relation
hold true at the input and output side of the variator, respectively, (h is the oil film thickness):
v21X ovx s21X v21Y ovy s21Y
¼ ¼ Cðse Þ; ¼ ¼ Cðse Þ ð45Þ
h oz se h oz se
v23X ovx s23X v23Y ovy s23Y
¼ ¼ Cðse Þ; ¼ ¼ Cðse Þ ð46Þ
h oz se h oz se
From the above written Eqs. (45) and (46) and from the Bair and Winer model (see Eq. (43)) it is
possible to obtain an explicit formulation for the shear stress acting on the rollers at the input
contact area:
   
6 v21X gjv j
 21 6 v21Y gjv j
 21
~s21X ¼ R~sL 1  e hsL ; ~s21Y ¼ R~sL 1  e hsL ð47Þ
p jv21 j p jv21 j
and at the output contact region:
   
6 v23X gjv j
 hs23 6 v23Y gjv j
 hs23
~s23X ¼ R~sL 1e L ; ~s23Y ¼ R~sL 1e L ð48Þ
p jv23 j p jv23 j
where ~s21X ¼ s21X K2 =FN , ~s21Y ¼ s21Y K2 =FN , ~s23X ¼ s23X K2 =FN , ~s23Y ¼ s23Y K2 =FN and
sL sL0 p p p~
~sL ¼ ¼ 0 þ a 0 ¼ ~sL0 þ a ð49Þ
E0 E E 6R
It is possible to rephrase Eqs. (47) and (48) in terms of the following dimensionless parameters:
v21X r21 a
~Y v21Y r21 ~
aX
¼ Crin  Y; ¼ X;
jx1 jr1 R  ~r1 jx1 jr1 R  ~r1
0 !2 !2 11=2 ð50Þ
jv21 j r21 ~
aY r21 ~
aX
¼ @ Crin  Y þ X A
jx1 jr1 R  ~r1 R  ~r1
934 G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942

v23X Crout r23 a


~Y v23Y r23 ~aX
¼  Y; ¼ X;
jx3 jr3 1  Crout R  ~r3 jx3 jr3 R  ~r3
0 !2 !2 11=2 ð51Þ
jv23 j Crout r23 ~aY r23 ~aX
¼@ þ Y þ X A
jx3 jr3 1  Crout R  ~r3 R  ~r3

21 j 23 j
Whereas, as regards, the quantities gjv hsL
and gjv
hsL
, that appear in Eqs. (47) and (48), the following
relations hold true:
0 !2 !2 11=2
gjv21 j g
~ r21 ~
aY r21 ~aX
¼ ~ 1 ð1 þ kÞ@ Crin 
x Y þ X A ð52Þ
hsL Hin~sL R  ~r1 R  ~r1
0 !2
gjv23 j g
~ Cr r ~
a
~ 1 ð1 þ kÞð1  Crin Þð1  Crout ÞsrID @
out 23 Y
¼ x þ Y
hsL Hout~sL 1  Crout R  ~r3
!2 11=2
r23 a
~X
þ X A ð53Þ
R  ~r3

where the dimensionless rotating velocity has been defined as:


g x1
~1 ¼ 0 0
x ð54Þ
E
This quantity takes into account the effect of the rotating speed of the input disc on the variator
behaviour. Hin ¼ h=qeqX in is the dimensionless thickness of the oil film at the input contact zone, it
can be evaluated by means of the hard EH lubrication formulas [32,33], where the above defined
dimensionless parameters are used:
h i0:67  0:134 
Hin ¼ 2:81 ð1 þ kÞð1  0:5Crin Þx~1 ~f0:53 R0:201 q~eqX 1  0:61  e0:73ein ð55Þ
in

In Eq. (55) the dimensionless pressure–viscosity coefficient ~f of the oil has been defined as (see
[32]):
 
~f ¼ fE0 ¼ Z1 ln g0 ð56Þ
~cp g1
As regards the dimensionless oil film Hout at the output contact, it can be calculated by means of
the relation:
Hout 2  Crout 1  0:61  e0:73eout
¼ ð1  Crin Þ0:67 s0:536
rID ð57Þ
Hin 2  Crin 1  0:61  e0:73ein
Equation (57) shows that, since the creep coefficients in a normal operative condition are suffi-
ciently small and the last term in Eq. (57), related to the ellipticity parameters ein and eout , is very
close to the unity, the ratio Hout =Hin can be evaluated by means of the simpler well approximate
relation
G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942 935

Hout
¼ s0:536
rID ð58Þ
Hin
that Eq. (58) shows that the oil film thickness at the input and output points of contact,
respectively, may differ significantly, especially at the extreme values of the speed ratio.

4.3. Calculation of the traction coefficient and spin momentum

The contact model described in Section 4, enable us to find an integral relation that allows for
the calculation of the traction coefficients lin and lout and the spin momentum coefficients vin and
vout , as reported in Eqs. (59) and (60).
Z 1 Z 2p
lin ¼ ~
aXin ~
aYin dR ~s21X R dw
0 0
Z 1 Z 2p ð59Þ
lout ¼ ~
aXout ~aYout dR ~s23X R dw
0 0
Z Z 2p  
~ aYin 1
aXin ~
vin ¼ dR aXin ~s21Y cos w  ~aYin ~s21X sin w R2 dw
~
R~r1 0 0
Z 1 Z 2p   ð60Þ
~
aXout ~
aYout
vout ¼ dR aXout ~s23Y cos w  ~aYout ~s23X sin w R2 dw
~
R~r3 0 0

where the following co-ordinate transformation has been applied:


X ¼ R cos w
; 0 6 R 6 1; 0 6 w 6 2p ð61Þ
Y ¼ R sin w

5. Results

Since the scope of the paper is to compare the half-toroidal and the full-toroidal traction drives,
it is necessary to specify which quantities will be kept constant during the calculations. These are
the ideal speed ratio srID , and the normal force FN at the points of contact, whereas the geometrical
quantities are reported in Table 1, and the fluid properties in Table 2. Fig. 8 shows the efficiency m
of the two variators as a function of the input dimensionless torque tin , for srID ¼ 1, and R ¼ 24.
This last parameter corresponds, for the given geometry to a maximum pressure value close to
2.2–2.3 GPa. Moreover, the angular velocity of the input disc has been chosen equal to
jx1 j ¼ 2000 [RPM], that corresponds to a dimensionless value x ~ 1 ¼ 2:95  1012 . As shown in
Fig. 8 the better efficiency of both variators is reached in the region of high tin . Moreover, the
efficiency of the full toroidal traction drive is smaller of about 4–5% points, compared to that of
the half-toroidal variator over the whole range of tin values. The reason of this difference is caused,
in the region of low tin , mostly by the worse torque efficiency of the full-toroidal variator (see Fig.
9). In fact, Fig. 10 shows that, because of the torque loss in the roller bearing and the larger
contact area (the radius of curvature ~r22 is bigger than in full-toroidal variator), the half-toroidal
936 G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942

Table 2
The fluid properties
Fluid properties:
T ¼ 99 °C
Absolute viscosity at the atmospheric pressure g0 ¼ 3:25  103 Pa s
Viscosity–pressure index Z1 ¼ 0:85
Pressure–viscosity coefficient f ¼ 1:71  108 Pa1
Limiting shear stress at atmospheric pressure sL0 ¼ 0:02  109 Pa
Limiting shear stress constant a ¼ 0:085
Pole pressure constant of Roelands viscosity model cp ¼ 1:96  108 Pa
Pole viscosity of Roekands viscosity model g1 ¼ 6:31  105 Pa s

ν 1

0.95

0.9

s r ID =1
0.85
ℜ = 24

0.8 Half-Toroidal
Full-Toroidal

0.75 2000 [RPM]

0.7
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
t in

Fig. 8. The efficiency of the variators as a function of the input traction coefficient tin .

1
ν torque

0.95

0.9

s r ID =1
0.85
ℜ = 24
0.8 Half-Toroidal
Full-Toroidal

0.75 2000 [RPM]

0.7
0 0.02 0.04 0.06 0.08 0.1
t in

Fig. 9. The torque efficiency of the variotors as a function of the input traction coefficient tin .
G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942 937

0.003
χ in

0.0025
s r ID =1

0.002
ℜ = 24
2000 [RPM]

0.0015

0.001 Half-Toroidal
Full-Toroidal
0.0005

0
0 0.02 0.04 0.06 0.08 0.1
t in

Fig. 10. The spin momentum of the variators as a function of the input traction coefficient tin .

traction drive is more effected by the spin momentum. But, because of its very small spin velocity
(see Fig. 5), the energy losses due to spin are less important than in the full-toroidal traction drive.
Fig. 9 shows that also for mid and high tin values the efficiency of full-toroidal CVT is smaller
than that of the half-toroidal traction drive, this is due mostly to the worse speed efficiency of the
full-toroidal variator as shown in Fig. 11. In fact, the large values of the spin motion, that affect
the full-toroidal CVT, cause the global sliding coefficient Cr to increase very fast as the requested
torque tout increases (see Fig. 12), thus causing larger power losses.
Regarding the influence of the dimensionless parameter R, Figs. 13 and 14 show how it affects
the efficiency of both CVT variators: the bigger R the higher the mechanical efficiency of the
variators. The explanation is simple: the efficiency of the variators is largely affected by the spin
losses, that in turn depends on the extension of the elliptical contact area. By reducing the load,
i.e. by incrementing R, the area of contact reduces its extension. This, in turn, causes a reduction

1
ν speed

0.95

0.9
s r ID =1
0.85
ℜ = 24
Half-Toroidal
0.8
Full-Toroidal
2000 [RPM]
0.75

0.7
0 0.02 0.04 0.06 0.08 0.1
t in

Fig. 11. The speed efficiency of the variators as a function of the input traction coefficient tin .
938 G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942

0.1
t out
0.09

0.08

0.07

0.06
s r ID =1
0.05
ℜ = 24
0.04
Half-Toroidal
0.03 Full-Toroidal
0.02 2000 [RPM]
0.01

0
0 0.05 0.1 0.15 0.2
Cr

Fig. 12. The traction capabilities of the variators: output traction coefficient tout as a function of the global sliding
coefficient Cr.

ν 1 Half-Toroidal

0.95

0.9 ℜ = 48
ℜ = 42
0.85 ℜ = 36
s r ID =1
0.8 ℜ = 30
2000 [RPM]

0.75
ℜ = 24

ℜ = 18
0.7
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
t in

Fig. 13. The efficiency of the half-toroidal variator as a function of the input traction parameter tin and for different
values of the load parameter R.

of the spin momentum, and therefore of the energy dissipated by the spin motion. Morevoer, since
the spin motion affects the FT variator more than the HF one, it is expected that the full-toroidal
variator be more sensible to the load variations than the half-toroidal CVT, as clearly demon-
strated in Figs. 13 and 14. Therefore, it may happen that for higher values of R the efficiency of
the full-toroidal variator overcomes that of the half-toroidal traction drive: see, for example, the
curves plotted for R ¼ 48 in Figs. 13 and 14. But, on the other hand, too big values of R reduce
the traction capability of the variators, i.e the maximum values of tin .
The simulations have shown that the parameter x ~ 1 does not affect appreciably the efficiency of
both variators. Moreover, it was shown that the FT traction drive, because of its particular
G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942 939

ν 1 Full-Toroidal

0.95 ℜ = 48
ℜ = 42
0.9
ℜ = 36

0.85
ℜ = 30

0.8 s r ID =1
ℜ = 24
2000 [RPM]
0.75

ℜ = 18
0.7
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
t in

Fig. 14. The efficiency of the full-toroidal variator as a function of the input traction parameter tin and for different
values of the load parameter R.

ν 1
Half-Toroidal

0.95

0.9

2000 [RPM]
0.85
ℜ = 24
s r ID =2/3
0.8 s r ID =1
s r ID =1.5
0.75

0.7
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
t in

Fig. 15. The efficiency of the half-toroidal variator as a function of the input traction parameter tin and for different
values of the speed ratio srID .

symmetry, is almost insensible to the actual value of ideal speed ratio srID , whereas the HT var-
iator results to be largely affected as shown in Fig. 15.
In conclusion, the analysis carried out, shows that the full-toroidal traction drive is unable to
achieve the same efficiency of the HT variator, despite the torque losses in the support bearing
that affect the latter. The main reason of this result is the very high spin motion of the FT variator
that produces high values of energy dissipation. It has been also shown that only two parameters
influence significantly the mechanical efficiency of the variators, these are the dimensionless load
parameter R and the dimensionless input torque tin .
940 G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942

6. Conclusions

The paper deals with the mechanical efficiency of full and half toroidal traction drives and has
the aim of comparing the performances of these two CVT typologies. A fully flooded isothermal
model of the contact between discs and rollers, based on the results of EHL theory, has been
implemented to evaluate the slip and spin losses. The analysis has shown that by optimizing the
roller geometry of the full toroidal variator, it is possible to reduce its spin momentum below the
value of the half-toroidal variator. But, since the energy dissipation due to the spin losses is
the product of the spin momentum and the spin velocity, the full-toroidal variator always results
to have a smaller efficiency because of its much bigger spin velocity. Moreover the full-toroidal
traction drive needs higher values of global creep (smaller speed efficiency) to transmit the same
torque of the HT traction drive. This causes an additional heating of the lubricant and a further
reduction of its traction capability and also of the mechanical efficiency. Moreover, the
mechanical efficiency of the half-toroidal CVT, in comparison to that of the full-toroidal variator,
is less affected by the value of the normal contact forces and often over the threshold of 90% on
the most part of the torque range. The full-toroidal CVT, instead, shows a different behavior, its
efficiency varies within a bigger range of values, and it is more affected by the normal load at the
contact between the rollers and discs.

References

[1] C. Brace, M. Deacon, N.D. Vaughan, R.W. Horrocks, C.R. Burrows, The compromise in reducing exhaust
emissions and fuel consumption from a Diesel CVT powertrain over typical usage cycles, in: Proceedings of
International Congress CVT’99, Eindhoven, The Netherlands, September 16–17, 1999, pp. 27–33.
[2] C. Brace, M. Deacon, N.D. Vaughan, C.R. Burrows, R.W. Horrocks, Integrated passenger car diesel CVT
powertrain control for economy and low emissions, in: ImechE International Seminar S540, Advanced Vehicle
Transmission and Powertrain Management, September 25–26, 1997.
[3] G. Carbone, L. Mangialardi, G. Mantriota, Fuel consumption of a mid class vehicle with infinitely variable
transmission, SAE Transaction 2001 Journal of Engines 110 (Section 3) (2002) 2474–2483.
[4] L. Mangialardi, G. Mantriota, The advantages of using continuously variable transmission in wind power systems,
Renewable Energy 2 (3) (1992) 201–209.
[5] L. Mangialardi, G. Mantriota, Continuously variable transmission with torque-sensing regulators in waterpum-
ping windmills, Renewable Energy 7 (2) (1992) 807–823.
[6] L. Mangialardi, G. Mantriota, Automatically regulated C.V.T. in wind power systems, Renewable Energy 4 (3)
(1994) 299–310.
[7] G. Carbone, L. Mangialardi, G. Mantriota, Influence of clearance between plates in metal pushing V-belt
dynamics, ASME Journal of Mechanical Design 124 (September) (2002) 543–557.
[8] G. Carbone, L. Mangialardi, G. Mantriota, EHL visco-plastic friction model in CVT shifting behaviour,
International Journal of Vehicle Design 32 (3–4) (2003) 332–357 (A special Issue on Advancements in the field of
vehicle transmission).
[9] G. Carbone, L. Mangialardi, G. Mantriota, The influence of pulley deformations on the shifting mechanism of
MVB-CVT, ASME Journal of Mechanical Design (in press).
[10] M. Patterson, the Full-Toroidal Variator in Theory and in Practice, Torotrak (Development) Ltd, in:
Proceedings of CVT ’96 Congress, Paper No. 9636394, SAE of Japan, Yokohama, September 11–12, 1996, pp.
95–100.
G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942 941

[11] M. Patterson, Traction drive contact optimisation, in: Proceedings of the Leeds-Lyon Symposium, Vehicle
Tribology, in: D. Dowson (Ed.), Tribology Series 18, Paper X (iv), Elsevier Science Publisher B.V., Amsterdam,
1991, pp. 295–300.
[12] H Machida, S Aihara, State of art of the traction drive CVT applied to automobiles, in: Proceedings of the Leeds-
Lyon Symposium 17, in: D. Dowson (Ed.), Tribology Series 18, Paper X (i), Elsevier Science Publisher B.V.,
Amsterdam, 1991, pp. 267–275.
[13] M Taniguchi, H Machida, S Aihara, Some experimental findings at EHL contacts under traction with a heavy-duty
traction drive, in: Proceedings of the Leeds-Lyon Symposium 17, in: D. Dowson (Ed.), Tribology Series 18, Paper
X (iii), Elsevier Science Publisher B.V., Amsterdam, 1991, pp. 287–294.
[14] H. Machida, Traction drive CVT up to date, in: Proceedings of the CVT’99 Congress, Eindhoven, The
Netherlands, September 16–17, 1999, pp. 71–76.
[15] H. Kumura, M. Nakano, T. Hibi, J. Sugihara, K. Kobayashi, Performance of a Dual-Cavity Half-Toroidal CVT
for Passenger Cars, in: Proceedings of the CVT’96 Congress, 1996.
[16] H. Kumura, J. Sugihara, Y. Partita, Y. Arakawa, M. Nakano, N. Maruyama, Development of a dual-cavity half-
toroidal CVT, in: Proceedings of the CVT’99 Congress, 1999, Eindhoven, The Netherlands, September 16–17,
1999, pp. 65–70.
[17] H. Machida, H. Itoh, T. Imanishi, H. Tanaka, Design principle of high power traction drive CVT, SAE Technical
Paper, in: International Congress and Exposition Detroit, No. 950675, Michigan, February 27–March 2, 1995,
pp.79–89.
[18] P. Tenberge, Ratio stability of toroidal traction drives, in: Proceedings of the CVT’99 Congress, inserire pag.,
Eindhoven, The Netherlands, September 16–17, 1999.
[19] H. Mori, T. Yamazaki, K. Kobayashi, T. Hibi, A study on the layout and ratio change characteristics of a dual-
cavity half-toroidal CVT, JSAE Review 22 (2001) 299–303.
[20] H. Tanaka, J. Yokohama, Aspect of variator control of half-toroidal CVT, in: Proceedings of the CVT 2002
Congress, 2002, pp. 23–33.
[21] H. Tanaka, Speed ratio control of a parallel layout double cavity half-toroidal CVT for four-wheel drive, JSAE
Review 23 (2002) 213–217.
[22] H. Tanaka, H. Machida, H. Hata, M. Nakano, Half-toroidal traction drive continuously variable power
transmission for automobiles, JSME Internal Journal, Series C 38 (4) (1995).
[23] H. Machida, T. Abe, Fatigue life analysis of a traction drive CVT, in: Proceedings of the CVT’96 Congress, 1996,
pp. 101–106.
[24] H. Tanaka, H. Machida, Half-toroidal traction-drive continuously variable power transmission, Proceedings of the
Institution of Mechanical Engineers 210 (July) (1996) 205–212.
[25] T. Yamamoto, K. Matsuda, T. Hibi, Analysis of the efficiency of a half-toroidal CVT, JSAE Review 22 (2001) 565–
570.
[26] H. Machida, Y. Murakami, Development of the POWERTOROS Unit half toroidal CVT, Motion and Control
No. 9 NSK, Oct. 2000, pp. 15–26.
[27] T. Imanishi, H. Machida, H. Tanaka, A geometric study of toroidal CVT, in: Proceedings of CVT ’96 Congress,
Paper No. 9636411, SAE of Japan, Yokohama, Sept. 11–12, 1996, pp. 107–111.
[28] T. Imanishi, H. Machida, Development of POWERTOROS Unit Half-Toroidal CVT (2), Motion and Control No.
10 NSK, April 2001, pp. 1–8.
[29] M. Raghavan, Kinematics of the full-toroidal traction drive variator, ASME Journal of Mechanical Design 124
(Sept.) (2002) 448–455.
[30] Y. Zhang, X. Zhang, W. Tobler, A systematic model for the analysis of contact, side slip and traction of toroidal
drives, ASME Journal of Mechanical Design 122 (December) (2000) 523–528.
[31] Z. Zou, Y. Zhang, X. Zhang, W. Tobler, Modeling and simulation of traction drive dynamics and control, ASME
Journal of Mechanical Design 123 (December) (2001) 556–561.
[32] B.J. Hamrock, Fundamentals of fluid film lubrication, Series in Mechanical Engineering, Book, McGraw-Hill,
1994 (ISBN 0-07-025956-9).
[33] D. Dowson, Modelling of elastohydrodynamic lubrication of real solids by real lubricants, Meccanica, Kluwer
Academic Publishers, vol. 33, 1998, pp. 47–58.
942 G. Carbone et al. / Mechanism and Machine Theory 39 (2004) 921–942

[34] K.L. Johnson, J.L. Tevaarwek, The influence of fluid rheology on the performance of traction drives, Journal of
Lubrication Technology, Transaction of the ASME 101 (July) (1979) 266–274.
[35] K.L. Johnson, J.L. Tevaarwek, Shear behaviour of elastohydrodynamic oil films, Proceedings of the Royal Society
of London A 356 (1977) 215–236.
[36] K.L. Johnson, R. Cameron, Shear behaviour of elastohydrodynamic oil films at high rolling contact pressures, in:
Proceedings of IMechE Tribology Group 182(Part 1), 1967–68.

You might also like