You are on page 1of 10

Lingyuan Kong Robert G.

Parker1
e-mail: parker.242@osu.edu Department of Mechanical Engineering, The Ohio State University, 206 W. 18th Avenue, Columbus, OH 43210

Mechanics of Serpentine Belt Drives with Tensioner Assemblies and Belt Bending Stiffness
Steady state analysis is conducted on a multipulley serpentine belt drive with a springloaded tensioner assembly. Classical creep theory is extended to incorporate belt bending stiffness as well as the belt stretching and centripetal accelerations. The belt is modeled as an axially moving EulerBernoulli beam with nonuniform speed due to belt extensibility and variation of belt tension. The geometry of the belt-pulley contact zones and the corresponding belt tension and friction distributions are the main factors affecting belt slip. Bending stiffness introduces nontrivial span deections, reduces the wrap angles, and makes the belt-pulley contact points unknown a priori. The free span boundary value problems (BVP) with undetermined boundaries are transformed to a xed boundary form. A two-loop iteration method, necessitated by the tensioner assembly, is developed to nd the system steady state. The effects of system parameters on serpentine drive behavior are explored in the context of an actual automotive belt drive. DOI: 10.1115/1.1903002

Introduction

Serpentine belt drives are the automotive and truck industry standard for front-end accessory drives. Through a long belt, accessories such as the alternator, air conditioner, power steering, and water pump are powered simultaneously from the crankshaft Fig. 1. Keeping proper belt tensions to provide adequate torque without belt slip is important for the durable, quiet operation of serpentine drives. For this reason, a tensioner assembly is introduced to automatically adjust to changing conditions. Accurate prediction of the system steady state under different operating conditions is helpful to designers to address competing engineering considerations subject to a variety of practical constraints on geometry and accessory loads. Engineers must balance factors like belt tension, potential for belt slip, bearing loads, belt fatigue, and noise in designing belt drives. The present analysis sets out a two-loop iteration procedure to determine the steady mechanics of general n-pulley serpentine drives with tensioner assemblies. The model considers both the belt spans and pulleys coupled by the friction and normal forces in the contact arcs. The belt model is an axially moving beam that is not restricted to small deection and curvature. Bending stiffness is an important theoretical and practical consideration because it leads to unknown belt-pulley contact points, curved free span deections, and nonuniform tension and speed distributions. Belt inertia is also modeled. Solution of the nonlinear equations yields the important quantities required in practice, including the nonuniform tension and speed distributions in the free spans and contact arcs, the friction forces generated in the sliding and adhesion contact zones, the belt-pulley contact points wrap angles, and the tensioner orientation angle. Other useful quantities such as maximum transmissible moment, power efciency, and a measure of belt-pulley vibration coupling can be computed from the derived solution. For a simplied three-pulley serpentine belt drive, Beikmann et al. 1 introduce a numerical method to calculate the system steady state where the belt is modeled as an axially moving string without bending stiffness and the belt speed is assumed constant everywhere. Kong and Parker 2 improve this model by consid1 Corresponding author. Contributed by the Mechanisms and Robotics Committee for publication in the JOURNAL OF MECHANICAL DESIGN. Manuscript received May 18, 2004; revised October 29, 2004. Associate Editor: J. S. Rastegar.

ering belt bending stiffness, but for each free span the belt-pulley contact points are the same as the string model, which are xed at the points of common tangency of the two bounding pulleys. From the calculated steady state, a belt-pulley coupling indicator is dened to quantify the dynamic coupling between the pulley rotations and the belt transverse vibrations. Both of these models assume that belt tensions are uniform in each free span. There is no analysis of the belt-pulley contact zones. Calculation of the system steady state with these restricted models is mainly for subsequent free vibration analysis. Belt-pulley mechanics has been investigated extensively, especially for a string model without bending stiffness 3,4. In spirit, these analyses focus on belt slip and tension/friction distribution. Creep theory 38 is typically used to describe the belt-pulley interaction in the contact zones where the belt is longitudinally extensible and the pulley is modeled as a rigid body. Friction from relative sliding between belt and pulley generates the accessory driving torque governed by a Coulomb law. The belt adheres to or slides over the pulley surfaces in distinct adhesion and sliding zones. For a symmetric two pulley drive symmetric means the driver and driven pulleys have the same radius, Gerbert 6 uses a string model and obtains the closed-form solution of the sliding/ adhesion zones as well as the tension and friction distributions along the belt. Bechtel et al. 7 and Rubin 8 improve this string model by incorporating belt stretching and centripetal accelerations. Bechtel et al. 7 investigate a symmetric two-pulley drive and give the difference between this model and common engineering models that neglect belt inertia/acceleration. Rubin 8 investigates an asymmetric two-pulley drive and highlights the inuence of stretching acceleration on drive performance criteria such as maximum transmissible moment and power efciency. By incorporating belt bending stiffness while retaining belt inertia and other essential modeling features, Kong and Parker 9 build on the model developed by Bechtel et al. and Rubin. It is shown that bending stiffness has important inuences on the system steady states. Neglecting it can result in signicant errors in circumstances of importance in applications. The steady state analysis in 9 is conducted on xed-center, two-pulley drives and cannot be directly applied to an n-pulley serpentine belt drive with a tensioner assembly. In this paper, the axially moving beam-pulley model is extended to n-pulley serpentine belt drives with tensioners such as that shown in Fig. 1. The tensioner assembly is a distinguishing feature that requires a new solution procedure. An iteration SEPTEMBER 2005, Vol. 127 / 957

Journal of Mechanical Design

Copyright 2005 by ASME

M = EI

Q = dM /ds .

Terms involving GV in Eq. 1 correspond to the stretching and centripetal accelerations. To complete the system, a constitutive law relating tension and speed is specied, T = EAm0V/G 1 , 3

Fig. 1 Seven pulley serpentine belt drive example dened in Table 1

method is proposed to nd the steady motion given the system geometric parameters, material properties, and operating conditions. The steady motions of an example automotive belt drive are examined to demonstrate design implications and underlying mechanics of practical systems.

where EA is the belt longitudinal stiffness and m0 is the belt mass per unit length in the rest state 710. Figure 2 presents a general single span stretched between two pulleys. One can view this as the ith span connecting the pulleys i and i + 1 in a serpentine belt drive, in which the pulleys and spans are numbered sequentially in the counterclockwise direction with the crankshaft being pulley 1. The beam in the ith free span is modeled as an Euler elastica. A local coordinate system is adopted where the local xi axis starts from the center of pulley i and goes through the center of pulley i + 1, and the y i axis points to the inside of the belt loop. Li is the distance between the centers of these two pulleys. Once the steady state is obtained in this local coordinate system, it is converted to the global coordinate system whose origin is the center of pulley 1 Fig. 1. Note if i = N, where N is the total number of pulleys, then this Nth span is between pulleys N and 1. Because there is no contact force in the free span, the governing equations in 1 become Ti GVi + EIii = 0 Ti GVii EIi = 0, 4 where the subscript i represents the ith span, and denotes differentiation with respect to the local arclength si. From the constitutive law 3, Ti GVi can be treated as one unknown eld variable pisi = Ti GVi = 1 G2 / m0EATi G2 / m0. Three boundary conditions can be written for the span in Fig. 2 T i 0 = T i0

Steady Solution for an Axially Moving Curved Beam

The steady motion of a single-span axially moving beam wrapping on two pulleys is a critical component of the iterative serpentine drive solution. Change in the thickness of the beam is neglected resulting in constant beam bending and longitudinal stiffness. Application of Newtons second law to an innitesimal segment of an axially moving beam and elimination of higher order terms in innitesimal quantities leads to the governing equations 9: dT GV d + EI = f ds ds T GV EI d 2 = n, ds2 1

i 0 =

1 Ri

i s i = L i =

1 . Ri+1

where s is the arc coordinate along the beam, Ts is the beam tension, G is the mass ow rate which is constant for steady state, Vs is the belt speed, EI is the beam bending rigidity, is the beam curvature, f s and ns are tangential and normal contact forces per unit length, respectively. The moment and shearing force are obtained from

Note that G and Ti0 are not initially specied. For reasons to be revealed later, here we assume that these two values are known. The problem has undetermined boundaries where the span endpoints and the total arclength of the beam, Li are calculated from the geometric requirements that the beam contacts and is tangent with both bounding pulleys. To solve the undetermined boundary problem, the set of governing equations in 4 is expanded by introducing three more eld variables isi, xisi, and y isi, where isi is the inclination angle measured from the positive xi direction and xisi, y isi are rectangular coordinates of the belt particles in the local coor-

Fig. 2 The ith span connecting the ith and i + 1th pulleys in a serpentine belt drive

958 / Vol. 127, SEPTEMBER 2005

Transactions of the ASME

dinate system Fig. 2. xisi and y isi are governed by xi = cos i and y i = sin i, as evident from the innitesimal belt segment in Fig. 2. isi is governed by i = i according to the conventional denition of the curvature. The contact and tangency conditions are xi02 + y i02 = Ri2 xi0 = Ri sin i0 xiLi Li2 + y iLi2 = Ri2+1 , xiLi Li = Ri+1 sin iLi . 6 7

The two equations in 6 guarantee that the two endpoints contact the surface of pulley i and i + 1, respectively. The rst equation in 7 further makes the belt tangent to pulley i, while the second equation in 7 enforces the belt tangency with pulley i + 1. Tangency is imposed by the angles on the pulleys to the contact points being equal to i0 and iLi, which are also the span endpoint inclination angles. i i = s i / L i, x Introduction of the nondimensional variables s i = y i / L i, i = p iL 2 i = L i i , p = x i / L i, y / EI into the governing equai tions and boundary conditions yields i + i = 0, i/ds id i/ds dp i 2 p i i/ds i = 0 d 2 i 1, 0s 8 i = i, di/ds i = cos i, i/ds dx i/ds i = sin i dy i 1, 0s 9 i0 = P0Li2/EI, p i 0 = L i/ R i, i1 = Li/Ri+1 , 10

Fig. 3

Free body diagram of tensioner assembly

additional differential equation and boundary condition suitably nondimensionlized: 1 dIisi = , dsi 1 + TisiEA 0 si Li with Ii0 = 0, 16

to the previous BVP formulation. Then, IiLi is equivalent to the desired integral term and is a natural product of the BVP solution.

3 Two-Loop Iteration Method for the System Steady State


A prominent characteristic of serpentine belt drives is the tensioner assembly, which consists of an idler pulley on a springloaded tensioner arm that rotates about the pivot O Figs. 1 and 3. Its purpose is to maintain proper belt tensions as the accessories are activated/deactivated or engine load/speed changes. The tensioner orientation is not specied a priori, instead its position is determined in the analysis. Because the above BVP formulation can only address spans between two xed-center pulleys or, more accurately, two pulleys with their center positions known, analysis of the spans adjacent to the tensioner pulley, such as spans 6 and 7 in Fig. 1, requires a specied tensioner angle . This necessitates an inner iteration loop to nd the tensioner angle within an outer iteration loop based on span tension. The outline follows the owchart in Fig. 4. For an initial guess of the span tension T10 see comment following 5, the steady solution for all spans is calculated for an initial guess of the tensioner orientation angle . Iteration on the angle for xed T10 continues until the moment balance for the tensioner and its adjacent spans is satised. Then, a compatibility condition requiring the back-calculated unstretched belt length to equal the specied input value is evaluated. If the error exceeds a tolerance, the span tension T10 is adjusted and the process repeats. 3.1 Outer Loop Iteration for a Multiple Pulley, FixedCenter Drive. Supposing the tensioner orientation angle is specied, then all the pulleys have xed centers. Two kinds of segments exist along the beam: free spans and contact zones. For the free spans, the BVP-solver based method presented in the previous section cannot be directly applied because neither G nor any of the boundary tensions Ti0 in 5 are specied or readily measured. Iteration is required. For the contact zones, the governing equation is the same as for the string model because bending stiffness only causes a constant moment uniformly distributed in the contact zones. For this assumed multiple pulley, xed-center drive, an iteration loop outer loop is used to nd the system steady state. This loop iterates on T10, the starting boundary tension for the rst span. For an assumed value of T10, the steady state of each segment is calculated sequentially around the drive along the counterclockwise direction as follows. Because the corresponding belt speed for the assumed T10 is given by V10 = R11, where 1 is the specied crankshaft pulley 1 rotational speed, G is obtained from the constitutive law 3. Thus, the steady state of the rst span is completely solvable SEPTEMBER 2005, Vol. 127 / 959

i02 + Li0y i02 = Ri2 , L i 0 x i12 = Ri2+1 , i 1 L i 2 + L i 1 y L i 1 x i0 = Ri sin i0, L i 0 x 11

i1 Li , 12 Ri+1 sin i1 = Lix

where the unknown nal span length Li is dened as a eld func i governed by tion Li = Lis i = 0, dLi/ds i 1. 0s 13

The reason to introduce the nondimensional variables and dene the unknown constant Li as an unknown function in Eq. 13 is to incorporate them into the standard form required for most generalpurpose BVP solvers, namely, ut = Ft, ut, gua, ub = 0, atb 14

boundary conditions,

where F, u, and g are n-dimensional vectors and F and g may be nonlinear. The standard form involves only coupled differential not algebraic equations for unknown functions ut dened on xed domain a,b. After these transformations, the problem is i 0 , 1. The cast entirely into the form 14 on the interval s seven boundary conditions 1012 equal the total order of the six differential Eqs. 8, 9, and 13. For an innitesimal segment of the belt with arclength dsi in Fig. 2, its unstretched length is dsi / 1 + i, where i = Tisi / EA is the longitudinal strain. The unstretched length of this whole span, required in subsequent steps, is Li0 =

Li

1 dsi . 1 + TisiEA

15

Direct integration can be used because the tension distribution Tisi is known from pisi = Tisi GVisi and the constitutive i law 3 or, one may dene Iisi = s 0 1 / 1 + Tisi / EAdsi and add the Journal of Mechanical Design

/R = n , Ti GV i i i

d Ti GV i

M i/ R i =

dsi*
iRi

= f i = in i, 17

f idsi* ,

where Ti and Vi are the belt tension and speed in the sliding zone of pulley i, i is the friction coefcient of pulley i, s* i is the local arclength along the sliding zone, and the applied torque M i is positive counterclockwise. The boundary tension T20 equals T1L1, and the belt tension and speed distribution are calculated using 17. Closed-form solution is readily available, with the tension and speed varying exponentially. The sliding angle 2 on pulley 2 is Fig. 2:

i =

1 Ti0 GVi0 , ln i Ti1Li1 GVi1Li1

i = 2.

18

For this sliding zone, the unstretched belt length is


0 L = i

iRi

1 1+ Tisi*EA

dsi* = 1

G 2/ m 0 R i EA i , i = 2. 19

ln

1 + EA/Ti1Li1 GVi1Li1 M i/Ri 1 + EA/Ti1Li1 GVi1Li1

In the adhesion zone, there is no friction because the belt and pulley have the same speed; the tension and speed are constant. T 2 2R 2 , Thus, the span 2 starting boundary tension T20 equals allowing calculation of the motion of the second span using the previously outlined single span analysis. The wrap angle 2 for pulley 2 is available from the known span 1 and 2 solutions that include the belt-pulley contact points. From geometry, the adhesion angle on pulley 2 is i = i i, i = 2. The unstretched length of the adhesion zone is
0 L = iRi/1 + Ti0EA, i

i = 2.

20

Fig. 4 Flowchart of iteration method involving two loops

by specifying i = 1 in Eqs. 813. For the contact zone on pulley 2, in which the belt shape is known and the curvature is constant, one only needs to nd the distributions of belt tension and friction. This contact zone can be divided into a sliding zone with belt-pulley slip and an adhesion zone with no relative motion. The adhesion zone precedes the sliding zone along the direction of belt motion Fig. 1 6. This is because for driven pulleys the friction force resists the belt motion, so the belt speed in the sliding zone is faster than that of the pulley surface. Belt tension and speed are constant in the adhesion zone; the tension increase along the contact arc in the direction of belt travel occurs entirely in the sliding zone. The increasing belt tension in the sliding zone along the belt motion direction leads to the corresponding increase of belt speed in the sliding zone because of the constitutive law 3. On the other hand, the belt speed in the adhesion zone is the same as that of the pulley surface, so even the slowest belt speed in the sliding zone is faster than that in the adhesion zone. If the sliding zone preceded the adhesion zone along the direction of belt motion for driven pulleys, then the belt speed would increase monotonically starting from a speed higher than that of the pulley surface and then reach the same speed as the pulley surface at the point where adhesion begins, which is an impossible situation. Similar reasoning applies to the driver pulley. For the sliding zone of pulley i, including i = 2, the governing equations are 960 / Vol. 127, SEPTEMBER 2005

This process repeats around the entire belt drive. There are several points to note in this solution procedure: a For any idler pulleys, such as the tensioner pulley, the torque is zero. The sliding angle is zero, and the wrap angle equals the adhesion angle. b The single span analysis outlined in the previous section is only suitable for free spans where both bounding pulleys contact the ribbed side of the belt. If one of the two bounding pulleys for the ith free span contacts the at side of the belt like pulleys 4 and 7 in Fig. 1, then some of the boundary conditions need modication while the differential equations remain the same. For example, for span 3 where pulley 4 is the end bounding pulley, the dimensional boundary conditions 5 and 7 must be changed to T 3 0 = T 30

3 0 =

1 R3

3 s 3 = L 3 =

1 , R4

21 22

x30 = R3 sin 30

R4 sin 3L3 = x3L3 L3 .

Span 6 is handled similarly. For spans where the starting bounding pulley contacts the at side of the belt like spans 4 and 7, the modications are the same in spirit and there is no difculty to nd the steady states of such free spans. c For the contact zones on pulley i, where i 2, the solution is obtained by replacing 2 with the individual pulley number i in Eqs. 1720. The only exception is for the sliding zone on pulley 1. In this case, the friction force exerted on the belt from the crankshaft is in the same direction as the belt motion, which is different from those of the other pulleys for which the friction Transactions of the ASME

Table 1 Physical properties of a practical serpentine belt drive Pulley No. Center X location Center Y location Radius R m Torque M N m Friction coefcient 0.95 0.95 0.95 0.95 0.95 0.95 0.95 L0 = 2.07918 m m0 = 0.107 kg/ m EA = 111,200 N In/Out property Ribbed Ribbed Ribbed Flat Ribbed Ribbed Flat

1 0 0 2 0.2116 m 0.0090 m 3 0.2317 m 0.1898 m 4 0.0796 m 0.2097 m 5 0.2026 m 0.2699 m 6 0.2000 m 0.1000 m 7 Crankshaft speed: Tensioner: 0 = 1.8997 rad Kt = 38.84 N m / rad Tensioner pivot Xo = 0.033 m Tensioner pivot Y o = 0.137 m

0.0970 m 24.82 0.0625 m 9.09 0.0291 m 0 0.04075 m 18.908 0.06685 m 2.382 0.08245 m 0 0.03775 m 1 = 2000 rpm clockwise Belt:

forces are opposite to the belt motion direction Fig. 1. This is because the crankshaft is the only driver pulley in the system while all other pulleys are driven ones. The torque exerted on pulley 1 is negative while all other torques are positive. For pulley 1, the rst of 17 and 20 still hold, but the remaining equations of 1719 become d T1 GV 1 ds* 1 = f 1 = 1n 1, M 1/ R 1 =

process repeated until convergence is achieved to within a chosen tolerance. 3.2 Inner Loop Iteration for Tensioner Orientation Angle. In the above iteration loop, it is assumed that the tensioner orientation angle is specied. In the analysis, however, the tensioner orientation angle is not known a priori because the tensioner assembly angle adjusts for different tensions and operating conditions. To determine the steady state of the entire system, an additional loop inner loop iterating on the tensioner orientation angle is added into the outer loop. The inner loop iterates on the tensioner and its adjacent spans/pulley to enforce tensioner moment balance. The process is outlined in Fig. 4. The tensioner orientation angle must satisfy the moment balance for the tensioner assembly about its pivot Fig. 3, where the strokes of the belt in the free spans are thickened. The balance of angular momentum with respect to the tensioner pivot applied to the control volume requires:

1R1

f 1ds* 1, 23

1 =

1 T10 GV10 , ln 1 T L GV L N N N N

0 = 1 L 1

G2/m0 R1 1 + EA/T10 GV10 + M 1/R1 ln . EA 1 1 + EA/T10 GV10 24

d For the analysis of the contact zone on pulley 1, the crankshaft torque M 1 is not specied the speed 1 is assumed specied. Instead, M 1 is calculated in the analysis. Integration of 1 along the entire belt loop leads to


fds =

1R1

f 1ds* 1

i=2

= TB GVBrTB QBrQB + TA GVArTA QArQA M S = 0. 28

iRi

f idsi* = 0.

25

Substitution of 17 and 23 into 25 yields M 1/ R 1 =

M /R .
i i=2

26

A criterion is needed to determine if the solution calculated for given T10 is the correct one. This iteration loop criterion is the compatibility condition

L
i=1

0 i

0 0 + L + L L0 = 0, i i

27
0

where L0 is the given total unstretched length of the belt, Li is the unstretched length of span i see Section 2 for the calculation 0 0 0 of Li , and L and L are unstretched lengths corresponding to i i the sliding and adhesion zones on pulley i. Physically, the compatibility Eq. 27 means that the unstretched belt length backcalculated from the steady state must equal the user-specied stress free belt length. For a practical serpentine belt drive, this user-specied value L0 is known by manufacturers or can be obtained by measurement in the stress free state. If the error in Eq. 27 is positive negative, T10 is increased decreased, and the Journal of Mechanical Design

The points A and B at the midpoints of the two adjacent spans are selected arbitrarily to dene the control volume. For the assumed tensioner orientation angle , the two adjacent spans are treated like those between xed-center pulleys, and there is no difculty to calculate the steady motions of these two spans. TA and TB are the tension forces at the midspan points while VA and VB are the corresponding speeds. rTB and rQB are the moment arm lengths with respect to the tensioner pivot O for the tension TB and shearing force QB at point B Fig. 3. rTA and rQA for the point A are not drawn in Fig. 3 for gure clarity. The shearing forces QA and QB and the moments M A and M B are calculated using 2 and the known free span solutions for the current iteration value of . The moment from the tensioner spring is M s = Ks o, where Ks is the tensioner pivot rotational stiffness and 0 is the specied tensioner orientation for zero moment in the pivot spring. If the error in Eq. 28 is positive negative, is increased decreased and the process repeated until convergence is achieved to within a chosen tolerance. In the above iteration process, the two loops repeat until both the compatibility condition 27 and the moment balance 28 are satised Fig. 4. One can use values calculated from the string model in Ref. 1 as initial guesses for T10 outer loop and inner loop to speed convergence. SEPTEMBER 2005, Vol. 127 / 961

Results and Discussion

Results are presented for a seven-pulley serpentine belt drive used in a production automobile Fig. 1. The physical properties are listed in Table 1. Note that the torque on pulley 1 crankshaft is calculated from 26 instead of being specied. The Ribbed/Flat property refers to whether the ribbed or at side of the belt contacts the pulley. Figure 5 depicts the system steady motions for three values of bending stiffness EI = 0.01, 0.1, and 0.3 N m2. Practical bending stiffness estimates for polyribbed belts used in serpentine belt drives fall roughly in the range EI = 0.01 0.1 N m2. Values may be much higher in other applications or in V belts. For small bending stiffness, the wrap angles are only slightly reduced; the span deections are close to those of the string model. Radial tickmarks indicate string model contact points on the lines of common tangency. For large bending stiffness, the deections are signicantly away from the straight line of the string model, especially for short spans or those with small tensions. The free span deections cause dynamic coupling between the pulley rotations and span transverse vibrations 2. Through this coupling mechanism, crankshaft speed uctuations can cause undesired belt span vibrations and noise. The steady states obtained in this study improve on those in 11 calculated from the xed belt-pulley contact points model 2 to better predict the coupled belt-pulley-tensioner natural frequencies, vibration modes, and dynamic response. The inuence of bending stiffness on the sliding angles are less pronounced than on the adhesion angles because the sliding angles are determined mainly by pulley torques independent of bending stiffness. Therefore, the dominant parts of the wrap angle reductions are in the adhesion zones. The adhesion zones function as power transmission reserve capacity. When more torque needs to be transmitted by an accessory pulley e.g., when the air conditioner is activated, a portion of the adhesion zone converts to sliding zone for that pulley. Bending stiffness greatly reduces the system capacity to transmit more power, or, in other words, bending stiffness makes the pulleys more prone to slip fully zero adhesion angle. Figures 6a6c show the numerical variations of these geometric angles with bending stiffness. The wrap angle reductions of small radius pulleys are much larger than those of large radius pulleys Fig. 6d. For example, for pulley 3 with the smallest radius, when bending stiffness increases from zero to 0.3 N m2, the wrap angle decreases by about 70 deg the adhesion angle is reduced to about 1 deg, at the edge of full slip. For pulley 1 with the largest radius, the wrap angle is reduced by about 16 deg. The phenomena observed in Fig. 6 can be explained by an approximate analytical solution. Supposing the bending stiffness effect is small, then the slopes in the free spans are low, and the wrap angle for pulley i is reduced by see 36 in the appendix

ireduced

1 Ri 1 Ri


EI + pi EI + pi EI , pN

EI , i pi1

1,

i = 1,

29

where the approximate pi calculated from the string model is used. The analytical and numerical solutions agree well for small bending stiffness as shown in Fig. 7. Equation 29 also explains why the wrap angle reduction for pulley 2 is less than that for pulley 5 in Fig. 6d even though pulley 2 has smaller radius. The two neighboring span tensions are higher for pulley 2 than those for pulley 5. To prevent full slip in systems where the bending stiffness effect is meaningful i.e., large bending stiffness, low tensions, or high speeds, accessories with large torques should have large pulley radii in addition to the common design requirement of large wrap angles for such accessories. This permits use of lower 962 / Vol. 127, SEPTEMBER 2005

Fig. 5 Steady state for the system properties specied in Table 1. a EI = 0.01; b EI = 0.1, c EI = 0.3 N m2. The xed beltpulley contact points for the string model are marked by short lines perpendicular to the pulleys. Dashed lines delineate the sliding and adhesion zones, denoted by A and S.

tensions, an important consideration in pulley bearing and belt life. Small pulley radii are sufcient for idler pulleys. From this perspective, an increased alternator pulley 3 radius in Fig. 5 is preferred because the alternator generally requires large torque. Figure 8 shows that the maximum span tensions increase sigTransactions of the ASME

nicantly, and almost linearly, with bending stiffness. Note tension and speed are not uniform along the belt free spans; the maximum tension appears around the span midpoint 9. Bending stiffness causes the belt to experience the cyclic stresses at a higher level. Neglecting bending stiffness results in overestimation of the belt fatigue life, especially for appreciable bending stiffness, which is contrary to safe design practice. The power efciency is dened as the ratio between the total power delivered to the driven pulleys and the power supplied by the driver pulley

M
i i=2

M 1 1

30

Power efciency changes little with bending stiffness. For the system specied in Table 1, varies less than 1% in a range near 99% over the bending stiffness values in Fig. 6. This results because the belt has large longitudinal stiffness EA, so tension variation along the belt causes only very small change in belt speed V = 1 + T / EAG / mo from the constitutive law 3. The nearly uniform belt speed makes the power efciency almost unity according to

i=2

M i i =

i=2

M iV i 0 / R i

M /R
i i=2

M 1 1

M 1V 1 0 / R 1

M 1/ R 1

= 1.

31

When the belt longitudinal stiffness is small, because the speed variation along the belt is more pronounced, the system power efciency is a much stronger function of longitudinal stiffness. For example, in Fig. 9 the belt longitudinal stiffness EA varies across one order of magnitude. The power efciency drops from about 99% to 94% with this decrease in longitudinal stiffness. Notice the steepness of the curve for the lower values of EA; further decrease dramatically decreases efciency. The longitudinal stiffness is dictated primarily by thin cords embedded in a rubber belt, so this property is at the designers discretion without affecting the belt adhesion properties. In practice, cords with a wide range of longitudinal stiffness are used in vehicle systems. Bending stiffness has little inuence on the tensioner orientation angle. The value of the tensioner orientation angle varies less than 1% about 165 deg over the bending stiffness range in Fig. 6. Tensioner orientation is similarly insensitive to friction coefcient and accessory loads, as shown later. This is also due to the large belt longitudinal stiffness EA. Usually the steady torque in the tensioner rotational spring under operating conditions is less than that in the rest state for the system in Fig. 1, it means that the tensioner orientation angle in the rest state is larger than under operating conditions. This is because, from the rest state to the operating states, more belt moves into the tensioner assembly part, whose two adjacent spans are the slackest ones of all spans. This reduces the tensions of these two spans, and less torque from the pivot rotational spring is needed to keep the moment balance 28. When EA is large, as is typical, a very small amount of belt moves into the tensioner assembly part under operating conditions, and only small changes in belt orientation angle are needed to accommodate the system requirements for the compatibility condition 27 and the moment balance 28. Therefore, for systems with large belt longitudinal stiffness EA, the tensioner orientation angle is a weak function of design parameters. In contrast, tensioner orientation depends strongly on longitudinal stiffness. Figure 9 shows that tensioner orientation angle decreases for decreasing belt longitudinal stiffness EA with a sharp slope for small EA. For comparison purpose, in Fig. 9, instead of having the same total unstrected belt length over the EA variation range, the tensioner orientation angles in the rest state are the same 167.51 deg. Under Journal of Mechanical Design

the same operating conditions, the variation in tensioner orientation angle from that in the rest state is higher for systems with small belt longitudinal stiffness EA. Figure 10 shows the effect of friction coefcient on the system steady states. The sliding and adhesion angles are sensitive to the friction coefcient, especially for the crankshaft, whose torque is much higher than the other accessories. As the friction coefcient decreases to 0.75, the crankshaft is at the edge of full slip. Belt wear reduces the friction coefcient, which affects the system capacity to function properly. Temperature also changes the friction coefcient. The friction coefcient has little inuence on the wrap angles, maximum span tensions, the tensioner orientation, and power efciency. This is also because of the large belt longitudinal stiffness EA. Conversion of the contact arcs between the sliding and adhesion zones due to the variation of friction coefcient causes only very small changes of belt tension/speed in the contact zones, which have little inuence on the rest of the system. Changing the torque of an individual pulley inuences only that pulley and any upstream components. For example, in Fig. 11, when the torque on pulley 5 is doubled, the sliding angles of pulleys 5 and 1 are increased with corresponding decrease in adhesion angles while the sliding angles of the pulleys between them decrease. The tensions of the spans between pulleys 1 and 5 increase. The rest of the system is almost unchanged. This is because the tensioner orientation angle is a weak function of accessory torques due to the large longitudinal stiffness EA as discussed earlier. Therefore, the tensions of the two spans adjacent to the tensioner remain almost the same from the moment balance about the tensioner assembly 28. This is the main design purpose of the tensioner assembly. Because the torque on pulley 6 is not changed, the sliding/adhesion angles on pulley 6 and the tension of span 5 remain almost the same. Because the accessory torque for pulley 5 is doubled, span 4 has higher tension. Part of the adhesion zone on this pulley converts to an enlarged sliding zone to transmit the increased torque. The increased tension of span 4 leads to higher tensions of the upstream spans 13. For middle pulleys like pulley 3, both of the two neighboring span tensions spans 3 and 2 increase. Thus, the beam tension throughout the pulley 3 sliding zone increases it increases exponentially from the tension of the neighboring slack span 3 to the tension of the neighboring tight span 2. The normal force between the belt and pulley also increases, and hence so does the friction force. Consequently, a smaller sliding zone is needed to transmit the unchanged torque M 3. For the crankshaft, because the torque is increased according to 26 and the tension of the neighboring span 7 is almost unchanged, the sliding zone increases.

Conclusions

Steady state analysis is conducted on a general multipulley serpentine belt drive with a tensioner assembly. The belt is modeled as an axially moving beam with bending stiffness, and the stretching and centripetal accelerations are incorporated. A two-loop iteration method is developed: the inner loop iterates on the tensioner orientation angle to satisfy the moment balance about the tensioner pivot while the outer loop iterates on the belt tension at the entry point of the crankshaft to satisfy a belt length compatibility condition. Key building blocks of the iteration are: a the solution for a single span wrapping on two bounding pulleys, a problem with undetermined boundaries solved by reformulating into a xed domain problem, and b the exact solution for the belt on the pulleys. The main conclusions include: 1. Bending stiffness signicantly decreases the wrap angles, making the pulleys more prone to slip fully due to reduction of the adhesion angles. 2. The maximum span tensions increase with bending stiffness, accelerating belt fatigue compared to commonly used engineering approximations based on string models. SEPTEMBER 2005, Vol. 127 / 963

Fig. 6 Variations of geometric angles for the system specied in Table 1: a wrap angles with bending stiffness; b adhesion angles with bending stiffness; c sliding angles with bending stiffness; d reduction of wrap angles 0 EI 0.3 N m2 with respect to pulley radii

Fig. 7 Comparison of variation of wrap angles with bending stiffness for the system specied in Table 1. numerical solution; --- approximate solution.

Fig. 8 Variations of span maximum tensions with bending stiffness for the system specied in Table 1

964 / Vol. 127, SEPTEMBER 2005

Transactions of the ASME

Fig. 9 Inuence of belt longitudinal stiffness on power efciency and tensioner orientation angle. The physical properties are from Table 1 with EI = 0.05 N m2.

3. Tensioner orientation and system power efciency are generally insensitive to changes of other design parameters because of the large belt longitudinal stiffness. These quantities are sensitive to belt longitudinal stiffness, where reducing longitudinal stiffness decreases the system power efciency and increases the variation of tensioner orientation angle for different operating conditions. 4. Friction coefcient has little inuence on wrap angles and the belt steady states in the free spans. The sliding and adhesion angles, however, are sensitive to changes of friction coefcient. Belt wear decreases friction coefcient, as can temperature, which reduces the system capacity to function properly with sufcient reserve power transmission capacity in the adhesion zones. 5. Changing the torque of an individual pulley inuences only the steady states of that pulley and upstream components with little effect on components in the downstream direction.

Fig. 10 Inuence of friction coefcient for the system specied in Table 1 with EI = 0.05 N m2

Acknowledgments
The authors thank Mark IV Automotive/Dayco Corporation and the National Science Foundation for support of this research.

Appendix
Suppose an axially moving belt is bounded between two pulleys with radius R1 and R2. When the bending stiffness is small, then the deection and curvature of the beam in the free spans are small. Elimination of higher order terms of innitesimal quantities in 1 leads to linearized equations for the free spans dT GV =0 ds Integrating these gives T GV = p = constant T GV EI d 2 = 0. ds2 32

s =

D + C1e p/EIs + C2e p/EIs , p 33

where p is the span tractive tension and D is a constant. For the stated assumptions, p is approximated well using the tension from a serpentine system string model 1 and uniform belt speed. Substitution of 0 = 0 = 1 / R1 and L = L = 1 / R2 gives indicates whether the pulley is above/below the horizontal line of common tangency Journal of Mechanical Design

Fig. 11 Inuence of accessory torques. The physical properties are from Table 1 with EI = 0.1 N m2 except that the torque on pulley 5 is doubled, as indicated

SEPTEMBER 2005, Vol. 127 / 965

1 D R1 p

1 EI e p/EIs R2 p

EI e p/EIsL . p

34

When EI 0, the inclination angle s should be zero recovering the solution for the string model, which gives D = 0. Furthermore, letting s = 0 or s = L in Eq. 34 gives the boundary inclination angles

1 R1

EI p 1 R

1 R2

EI . p

35

The total wrap angle around a pulley is reduced by

reduced


EI + p1

EI , p2

36

where p1 and p2 are the tractive tensions of the two neighboring spans.

References
1 Beikmann, R. S., Perkins, N. C., and Ulsoy, A. G., 1996, Design and Analysis of Automotive Serpentine Belt Drive Systems for Steady State Performance,

ASME J. Mech. Des., 119, pp. 162168. 2 Kong, L., and Parker, R. G., 2003, Equilibrium and Belt-Pulley Vibration Coupling in Serpentine Belt Drives, ASME J. Appl. Mech., 705, pp. 739 750. 3 Fawcett, J. N., 1981, Chain and Belt Drives: A Review, Shock and Vibration Dig., 13, pp. 512. 4 Johnson, K. L., 1985, Contact Mechanics, Cambridge University Press. 5 Gerbert, G., and Sorge, F., 2002, Full Sliding Adhesive-Like Contact of V-Belts, ASME J. Mech. Des., 1244, pp. 706712. 6 Gerbert, G., 1999, Traction Belt Mechanics, Chalmers University of Technology, Sweden. 7 Bechtel, S. E., Vohra, S., Jacob, K. I., and Carlson, C. D., 2000, The Stretching and Slipping of Belts and Fibers on Pulleys, ASME J. Appl. Mech., 67, pp. 197206. 8 Rubin, M. B., 2000, An Exact Solution for Steady Motion of an Extensible Belt in Multipulley Belt Drive Systems, ASME J. Mech. Des., 122, pp. 311316. 9 Kong, L., and Parker, R. G., 2004, Steady State Mechanics of Belt-Pulley Systems, ASME J. Appl. Mech., 721, pp. 2534. 10 Leamy, M. J., 2004, On a New Perturbation Method for the Analysis of Unsteady Belt-Drive Operation, ASME J. Appl. Mech., in press. 11 Kong, L., and Parker, R. G., 2004, Coupled Belt-Pulley Vibration in Serpentine Drives with Belt Bending Stiffness, ASME J. Appl. Mech., 711, pp. 109119.

966 / Vol. 127, SEPTEMBER 2005

Transactions of the ASME

You might also like