Sei sulla pagina 1di 36

Published in Proceedings of the 26th Forging Industry Technical Conference, Chicago, IL, November 2005

A LITERATURE REVIEW ON DURABILITY EVALUATION OF CRANKSHAFTS INCLUDING COMPARISONS OF COMPETING MANUFACTURING PROCESSES AND COST ANALYSIS
M. ZOROUFI1 AND A. FATEMI2 FORMERLY POSTDOCTORAL RESEARCH ASSOCIATE, 2PROFESSOR MECHANICAL, INDUSTRIAL AND MANUFACTURING ENGINEERING DEPARTMENT THE UNIVERSITY OF TOLEDO, TOLEDO, OHIO
1

ABSTRACT This paper presents the result of a literature survey focused on fatigue performance evaluation and comparisons of forged steel and cast iron crankshafts. Crankshaft specifications, operating conditions, and various failure sources are first reviewed. Then design aspects and manufacturing procedures for crankshafts are discussed. This includes a review of the effects of influential parameters such as residual stresses on fatigue behavior and methods of their production in crankshafts. The common crankshaft material and manufacturing process technologies currently in use are then compared with regards to their durability performance. This is followed by a discussion of durability assessment procedures used for crankshafts, as well as bench testing and experimental techniques. Geometric optimization of crankshafts is also briefly discussed and cost analysis and potential cost saving potentials from several studies in the literature are presented.

INTRODUCTION Crankshaft is among the largest components in internal combustion engines. It is employed in different types of engines, from small one cylinder lawn-mowers to large multi-cylinder diesel engines. Crankshaft is one of the most critically loaded components and experiences cyclic loads in the form of bending and torsion during its service life. Figure 1 shows a typical automotive crankshaft consisting of main journals (located on shaft centerline), connecting rod journals (located off-centerline), a trust bearing journal, cranks that connect the journals and hold the component together, and a number of auxiliary parts. A typical automotive crankshaft produced by hot impression forging of vanadium microalloyed steel for V6 engine is about 500 mm in length and 25 kg in weight. Such steel has a tensile strength of about 800 MPa, yield strength of about 500 MPa, and fatigue strength of about 375 MPa (1). Fillets of crank pins (rod journals) are the critical locations of the crankshaft that endure the highest level of stress under service loading. Figure 2 shows an automotive ductile iron crankshaft and the fillet of the crank pin. Two main sources of loading exist for crankshafts; gas load due to the combustion process in the combustion chamber

transmitted to the crankshaft by connecting rods, and inertia load due to the mass of the component and its attachments. Due to the dynamic nature of the system, both the gas and the inertia forces apply bending and torque to the component. The bending moments are reacted by journal bearings. Figure 3 shows the variation of bending and torque with respect to crank angle, separately measured at the fillet of an eight cylinder engine crankshaft (2). Identifying sources of failure of crankshaft is critical to its design and optimization. On the crankshaft surface, fillets behave as stress raisers and cracks can nucleate at their surface due to combined cyclic bending and torsion and grow inward. Silva (3) separates failure sources into three categories; operating sources, mechanical sources, and repairing sources. The operating sources include oil absence, defective lubrication on journals, high operating oil temperature, and improper use of the engine. Mechanical sources include misalignments of the crankshaft on assembly, improper journal bearings (wrong size), lack of control on the clearance between journals and bearings, and crankshaft vibration. Repairing sources include misalignments of the journals due to improper grinding, misalignments of the crankshaft, high stress concentrations due to improper grinding at the radius on either side of the journals, high surface roughness due to improper grinding or wear, improper welding or nitriding, and straightening operations. In this review paper, first the conceptual design process for crankshafts and the tasks involved are discussed. Then, manufacturing process steps for cast and forged crankshafts are presented and the influence of residual stresses on fatigue behavior from fillet rolling, shot peening, nitriding, and induction hardening are discussed. This is followed by a comparison of crankshaft material and manufacturing process technologies currently in use in terms of their durability performance. Several durability assessment procedures used for crankshafts are then presented, including discussions of fatigue crack growth, geometry effects, and surface hardening. Bench testing and experimental techniques to evaluate fatigue performance of crankshafts are also presented. Next, geometric optimization of crankshafts is briefly discussed, and finally, cost analysis and potential cost saving potentials from several studies in the literature are presented.

DESIGN AND MANUFACTURING CONSIDERATIONS A thorough conceptual design process for a crankshaft requires input design data from the engine specifications and operating conditions, design task including design for rigidity, static strength and durability, and manufacturing processes and considerations. Figure 4 shows a chart of crankshaft design tasks prepared by Dubensky (4). According to this flowchart, preliminary dimensions are specified based on the engine design data and previously designed comparable components. This preliminary design should be verified for rigidity, deformations, static strength, and fatigue strength under different load-case scenarios (i.e. bending, torsion, and combined bending and torsion loading conditions) considering appropriate factors of safety. Other design factors such as

lubrication requirements, frequencies of vibration in torsion and bending and engine sound level are to be included afterwards. When the basic requirements of the preliminary design are fulfilled, alternative manufacturing processes should be evaluated to obtain the most feasible and cost effective manufacturing plan. Following preparation of manufacturing plan and simulating the processes, a prototype component should be manufactured and tested to verify the design requirements. A detailed flowchart including a step-by-step calculation procedure for crankshaft conceptual design is provided by Dubensky (4). Crankshafts are typically manufactured by casting and forging processes. Figure 5 shows the crankshaft manufacturing process flow for sand casting and forging. Manufacturing by forging has the advantage of obtaining a homogeneous part that exhibits less number of microstructural voids and defects compared to casting. In addition, directional properties resulting from the forging process help the part acquire higher toughness and strength in the grain-flow direction. While designing the forging process for crankshaft, the grain-flow direction can be aligned with the direction of maximum stress that is applied to the component (along the axis of the shaft and related to bending). According to Shamasunder (1), there are three typical stages in crankshaft forging; reducer rolling (to prepare a preform), blocker forging (to give a basic profile to the preform), and finisher forging (to give the desired contours to the crankshaft). In each stage, a well-planned deformation is induced to ensure metal flow into the die cavity in both top and bottom dies. As the forging continues, top and bottom dies squeeze the billet. Each point in the workpiece moves in a particular direction with a specific velocity as determined by the die cavity profiles. Metal flow pattern will result in a complete filling of the die cavity to produce a sound forging quality. Grain flow pattern is an indicator of forging quality as well as directional strength. Manufacturing of the blank forged crankshaft could be followed by a number of post-processing steps including machining, heat treatment, induction hardening, and surface rolling. The heat treatment step is not applied if microalloyed steels are selected as materials of construction for the crankshaft, lending to considerable cost saving. A typical procedure of the post-processing stage of a 42CrMo4 alloy steel crankshaft including machining, heat treatment and surface hardening is shown in Figure 6. According to this procedure (5), the forging process is followed by annealing to remove the unwanted residual stresses generated by the refinement procedure applied to forgings. The annealed parts are subject to straightening after quenching and tempering, to correct for deformations due to the heat treatment. Additional annealing is then performed to remove the residual stresses generated during straightening. Next, in the machining stage, the part is turned and ground to obtain the required dimensions and tolerances. Finish grinding should be carefully selected to avoid the occurrence of unfavorable tensile residual stress distributions that would remain in the material even after induction surface hardening and grinding and would reduce fatigue strength of the material. Induction surface hardening is followed by stress annealing if the depth of the surface hardening is smaller than the depth of the damaged layer, since in this way it is possible to change the unfavorable stress state in the surface layer induced by

machining. After induction surface hardening, the magnitude and distribution of residual stresses should be such that they contribute to the fatigue strength of the material. Induction surface hardening is followed by finish grinding and non-destructive magnetic inspection of the surface to reveal the possible existence of cracks. Fillet rolling has traditionally been used to induce compressive residual stresses at the crankshaft fillets. The compressive residual stress generated at this critical area increases fatigue life of the component. Optical measurements at the deep-rolled crankshaft fillets show that the amount of compressive residual stress increases significantly in the axial direction, which coincides with the direction of bending stresses the component is subject to (6). In addition, prior to rolling crankshaft fillets, the stress gradients are very high near the fillet surface such that the stress magnitude drops very quickly with increasing distance away from the fillet surface (7). This makes the process of fillet rolling quite beneficial. The influence of the residual stresses induced by the fillet rolling process on the fatigue process of a ductile cast iron crankshaft section under bending was studied by Chien et al. (7) using the fracture mechanics approach. They investigated fillet rolling process based on the shadowgraphs of the fillet surface profiles before and after the rolling process in an elasticplastic finite element analysis with consideration of the kinematic hardening rule. A linear elastic fracture mechanics approach was employed to understand the fatigue crack propagation process by investigating the stress intensity factors of cracks initiating from the fillet surface. An effective stress intensity factor, which combines the stress intensity factors due to the bending moment and due to the residual stress, was defined. The effective stress intensity factor range was then approximated and compared to an assumed threshold stress intensity factor range to determine if the crack can continue to propagate for a given crack length. To validate the computational results, resonant bending fatigue tests were performed and fourbubble failure criterion was employed. According to this criterion, oil is applied to the fillet surface during the resonant bending fatigue test. For a given bending moment, the crankshaft fatigue life is determined when four pinhead-sized bubbles within a 6.35 mm area appear on the fillet surface. At this point, the test is suspended. It was found that this failure criterion that is based on cracks nucleated in the surface of the fillet is too conservative when the residual stress near the fillet due to rolling and the high stress concentration near the fillet are accounted for. Park et al. (8) studied the effect of nitriding and fillet rolling on fatigue life of crankshafts made of microalloyed steel and quenched and tempered alloy steel. Figure 7 shows component test results of this study for various surface treatments, i.e. nitriding, fillet rolling (500 kgf and 900 kgf) and bare samples. These experiments indicate that surface modification can increase endurance limit of crankshaft significantly. Both fillet rolled (at 900 kgf of rolling) and nitrided samples show more than a factor of 1.8 increase in fatigue limit. In addition, it was found that higher rolling force induces higher compressive residual stress on the crank surface, leading to better endurance limit characteristics. Nevertheless, with residual stress reaching a certain optimum level, additional load only results in plastic deformation that is detrimental to fatigue limit. This optimum residual stress could be found experimentally or numerically

using FEA. It should be noted that Gligorijevic et al. (9) reported a factor of 1.3 improvement in fatigue strength of nodular cast iron diesel engine crankshafts after nitriding. Gundlach et al. (10) investigated the influence of fillet rolling and shot-peening in austempered ductile iron that is commonly used as an alternative for forged steel in manufacturing crankshafts and camshafts. They performed rotating bending fatigue tests on smooth and notched hour-glass shaped specimens shown in Figure 8, made from the cylindrical end of the crankshaft. Fillet specimens had tangential fillets of 1.02 mm, 1.52 mm, or 3.05 mm radius. They found that fillet rolling improves the fatigue strength of all three fillet geometries considered. They suggest that the fillet rolling force be increased as the fillet radius increases to attain the same degree of compressive stress and subsequently obtain comparable improvements in fatigue strength. Using optimum conditions for fillet rolling will improve the fatigue strength of the filleted specimens equal to or even above that of the smooth specimens. They also found that shot peening also improves the fatigue strength of the filleted specimens, but to a lesser degree than fillet rolling. Although the compressive residual surface stress by shot peening was higher than that of the fillet rolled specimens, the depth of penetration of the compressive stress developed by shot peening was appreciably shallower than by fillet rolling. Figure 9 shows the improvement of fatigue strength of the ADI specimens by fillet rolling or shot peening. However, they emphasized that both shot peening and fillet rolling decrease the notch sensitivity of the ADI and move the fracture from the fillets to the gage section of the test specimens. Hoffmann and Turonek (11) examined forged steel crankshafts in terms of their mechanical properties and fillet rolling. The general material grades they tested including heat treatments and hardness are listed in Table 1. Sections of the crankshafts were tested to generate complete reverse bending fatigue data. Table 2 lists the fatigue data for the steel grades developed separately. The percentage increase in fatigue strength attributed to fillet rolling for different materials are 23% for carbon steel (CS) and high strength steel (HS), 44% for microalloyed steel MA1 with 8% V, and 25% for microalloyed steel MA2 with 6% V. The mechanical property data are summarized in Table 3. Induction surface hardening is also used to induce compressive residual stresses at the critical locations of the crankshaft. A study by Grum (5) shows that significant favorable residual stress ranging from 1020 to 1060 MPa at a depth of around 250 m, then slowly dropping to a depth of 3.5 mm to around 800 MPa, is generated at bearing locations by induction hardening of a forged CrMo steel crankshaft. Therefore, by gently varying the hardness and through compressive residual stresses in the transition area, it is possible to reduce the detrimental effect of the notch induced by stress concentration in parts under cyclic loading. However, according to the same study, a major difficulty in induction surface hardening is to ensure a very slight variation in hardness and the existence of compressive residual stresses in transition areas to the hardness of the base material.

PERFORMANCE COMPARISON OF COMPETING MANUFACTURING TECHNIQUES The major crankshaft material and manufacturing processes currently in use are forged steel, nodular cast iron and austempered ductile iron (ADI). Chatterley and Murrell (12) compared fatigue strength of crankshafts made of forged steel, ductile iron and austempered ductile iron by conducting bending fatigue tests. Tests were performed on nitrided 1% chromium/molybdenum forged steel, fillet-rolled ductile iron with 700 MPa tensile strength and 2% elongation, and fillet-rolled ADI (the austempered version of the ductile iron used). All crankshafts were designed to operate within a 4cylinder turbo-charged diesel engine. The fatigue tests were carried out on a constant amplitude mechanical stroking machine. The crankshafts were firmly clamped with split clamping blocks across two main bearing journals, with the bending moment being applied by means of a moment arm bolted onto an adapter press-fitted onto either the front (nose) or flywheel end of the shafts. Results of the fatigue tests are summarized in 4. Fatigue strength at ten million cycles of the rolled ADI crankshaft was found to be inferior to the forged steel crankshafts. However, the ADI showed better fatigue strength than ductile iron. Pichard et al. (13) investigated the possibility of replacing the traditional alloyed steels and cast iron with control-cooled microalloyed steel in order to save cost and increase productivity without compromising the mechanical properties including fatigue. Based on the results of their research, 35MV7 control-cooled microalloyed steel shows similar tensile and rotating bending fatigue strength as AISI 4142 quenched and tempered steel, while an improved machinable version of 35MV7 shows 40% higher turning index* and 160% higher drilling index compared to AISI 4142 quenched and tempered steel. In addition, the behavior of 35MV7 microalloyed steel after induction hardening is equivalent to that of the referenced quenched and tempered steel. For the response to ion nitriding compared to the nitrided heat treated steels such as AISI 4142 and AISI 1042 grades, the surface values are identical, the conventional nitrided depth is greater and the treated layers are more homogeneous for the 35MV7 steel. For short nitriding treatments, the results obtained on the 35MV7 steel are much better than those obtained for AISI 1042 quenched and tempered steel, the fatigue limit increased by about 135%. In comparison with a quenched and tempered highly alloyed steel used in crankshafts (32CDV13) and for short nitriding treatments, the fatigue limit of 35MV7 steel is only 10% lower, while significant cost saving could be made by using 35MV7 steel. When forged steel are compared to cast iron and alloyed ductile iron used in crankshafts, the fatigue properties of forged steels are better that that of cast iron. Table 5 shows fatigue experiment results on ductile iron, alloyed ductile iron, quenched and tempered steel and microalloyed steel specimens made of crankshafts. A 13% cost reduction was obtained for the final component by replacing the traditional AISI 4142 steel with 35MV7 control-cooled microalloyed steel. This includes 10% savings on the

Cutting speed that causes flank wear of 0.2 mm on the tip of the tool in 20 min continuous machining. Cutting speed at which it is just possible to bore 50 holes in a plate 40 mm thick before the tool becomes unusable.

unfinished piece, 15% saving on mechanical operations and 15% saving on ion nitriding treatment. Gligorijevic et al. (9) named a number of advantages of nodular cast iron in crankshaft applications. According to them, nodular cast irons combine the favorable characteristics of other ductile materials, such as steel, with other advantages, such as easy machinability, design flexibility (i.e. the free selection of shape of the component and thus the ability to integrate several functions in a single part), high dimensional accuracy of the raw castings and the resultant cost reduction in final machining. They state that nodular cast iron has higher damping capacity and lower notch sensitivity. However, these statements may have overlooked the generally inferior fatigue properties of nodular cast iron compared to steels, and the significant role of casting voids and defects in reducing the crankshaft fatigue strength. In the study of Gundlach et al. (10) discussed earlier, notch sensitivity of an ADI, namely a grade between ASTM A897 Grade 1 and 2, was investigated. As the severity of the fillet radius increased from 3.0 to 1.5 to 1.0 mm, the fatigue strength at ten million cycles decreased from 335 MPa to 282 MPa to 253 MPa. Figure 9 shows the comparison of fatigue strengths of smooth and filleted specimens. The results of this study do not agree with those of Chatterley and Murrell (12) discussed earlier with respect to the effect of fillet rolling on durability performance of ADI.

DURABILITY ASSESSMENT OF CRANKSHAFTS Durability assessment of crankshafts includes material and component testing, stress and strain analysis, and fatigue or fracture analysis. Material testing includes hardness, monotonic, cyclic, impact, fatigue and fracture (crack growth) tests on specimens made from the component or from the base material used in manufacturing the component. Component testing includes fatigue tests under bending, torsion, or combined bending-torsion loading conditions. Dynamic stress and strain analysis must be conducted due to the nature of the loading applied to the component. Complexity of the geometry of most crankshafts necessitates employing finite element analysis tools. Figure 10 illustrated the stress distribution obtained by static analysis due to gas load in the fillet of a crankshaft. Although static analysis is commonly used in designing crankshafts, the nature of loading especially the torsional vibrations makes transient stress analysis inevitable. The fillet has shown to be the most critical location under primary static loading in crankshafts. Nevertheless, performing transient analysis on a three dimensional solid model of a crankshaft is costly and time consuming. Payer et al. (14) developed a two-step technique to perform nonlinear transient analysis of crankshafts combining a beam-mass model and a solid element model. Using FEA, two major steps are used to calculate the transient stress behavior of the crankshaft; the first step is the calculation of time dependent deformations by a step-bystep integration. Using a rotating beam-mass-model of the crankshaft, a time dependent nonlinear oil film model and a model of the main bearing wall structure, the mass,

damping, and stiffness matrices are built at each time step. The system of resulting equations is then solved by an iterative technique. In the second step, those transient deformations are enforced to a solid element model of the crankshaft to determine its time dependent stress behavior. The major advantage of using the two steps is reduction of CPU time for calculations. This is because the number of degrees of freedom for performing of step one is low and therefore enables an efficient solution. Furthermore, the stiffness matrix of the solid element model for step two needs only to be built up once. In order to estimate fatigue life of crankshafts, Prakash et al. (15) performed stress and fatigue analysis on three example parts, belonging to three different classes of engines. The classical method of crankshaft stress analysis (by representing crankshaft as a series of rigid disks separated by stiff weightless shafts) and an FEM based approach using ANSYS code were employed to obtain natural frequencies, critical modes and critical speeds, and amplitudes and stresses in the critical modes. A fatigue analysis was also performed and the effect of variation of fatigue properties of the material on failure of the parts was investigated. Henry et al. (16) presented a procedure to assess crankshaft durability. This procedure as shown in Figure 11 consists of four main steps. The first step is modeling and load preparation that includes mesh generation, calculation of internal static loads (mass), external loads (gas and inertia) and torsional dynamic response due to rotation. The second step is the finite element method calculation including generating input files for separate loading conditions. Third step is the boundary condition file generation. The final step involves the fatigue safety factor determination. This procedure was implemented for a nodular cast iron diesel engine crankshaft. The initial (before optimization) eight counterweight crankshaft design is illustrated in Figure 12. The worst load condition with respect to fatigue safety factor was at 3500 rpm engine speed. At this speed one of the fillets has the lowest fatigue safety factor. The variation of calculated torsional moment and bearing load with respect to engine speed is shown in Figure 13. Using the operating conditions of Figure 13, the stress variation for the most critical location (crank pin fillet) is obtained as shown in Figure 14. Fatigue crack growth analysis of a diesel engine forged steel crankshaft was investigated by Guagliano and Vergani (17) and Guagliano et al. (18). They experimentally showed that with geometry like the crankshaft, the crack grows faster on the free surface while the central part of the crack front becomes straighter. Based on this observation, two methods were compared; the first considers a three dimensional model with a crack modeled over its profile from the internal depth to the external surface. In order to determine the stress intensity factors concerning modes I and II a very fine mesh near the crack tip is required which involves a large number of nodes and elements, and a large computational time. The second approach uses two dimensional models with a straight crack front and with the depth of the real crack, offering simpler models and less computational time. Comparison of the two different approaches based on KI and KII values show good agreement. In addition, the model was validated using strain-gage measurements during tests. Figure 15 shows the axial strain pattern as a result of the two-dimensional plain analysis and its comparison with

experimental results obtained by strain-gage measurements. In addition, they observed a linear empirical relation between the crack depth and its external dimension. Regulskii et al. (19) investigated the effect of geometry and surface hardening on fatigue strength of crankshafts of a two-cylinder motorcycle engine made of VCh 50-2 high-strength cast iron. Non-hardened crankshafts with a fillet, and hardened and nonhardened crankshafts without a fillet were analyzed. Figure 16 shows the results of fatigue testing of the crankshafts. The life of the crankshafts subjected to vibration surface hardening appears to be, on the average, 8% higher than that of the nonhardened crankshafts with a fillet. This gain is insignificant when it comes to cyclic life and attests to the low efficiency of the vibration surface hardening in this case, mostly because the scatter in the lives of these crankshafts is much higher than that of the crankshafts with a fillet. The decrease in the stress concentration due to the 2 mm radius fillet caused an increase in life up to the level comparable to that provided by hardening the shafts without a fillet. The crack developed along the circumference of the crankpin and, upon reaching the length of the order of 0.25 - 0.30 of its perimeter, it moved to the central cheek and propagated in the direction normal to the crank plane. In a number of cases, the occurrence of the first crack was shortly followed by the occurrence of the second one initiated in a similar manner and then a competitive development of the two cracks took place. The decrease in the endurance of the crankshafts tested compared to smooth specimens, as calculated from the ratio of smooth specimen fatigue limit to crankshaft fatigue limit was found to be in the range of 3.0 to 4.3, due to the effect of geometry, treatment procedure and other technological factors. EXPERIMENTAL TECHNIQUES AND BENCH TESTING Component testing is considered as a required step in durability assessment and its results incorporate the effects of geometry, surface finish, residual stresses, and directionality of properties on fatigue behavior. Because of the high cost of the component and test systems, and complex geometry of multiple-bearing one-piece crankshafts, fractions of the crankshafts consisting of one crank with two coaxially located main journals fixed in the grips of the testing machine are commonly used for fatigue testing. Although this fractionizing somewhat deviates from the real component service conditions, it provides a considerable reduction in the number of specimens and the cost of tests. Yu et al. (20) and Chien et al. (7) performed resonant bending fatigue tests on SAE J434C D5506 cast iron crankshaft sections. In the first study, resonant frequencies of a resonant bending system with notched crankshaft sections were obtained experimentally and numerically in order to investigate the effect of notch depth on the change of resonant frequency of the system. Resonant frequencies of the resonant bending system with crankshaft sections were obtained before and after introduction of the notches and the frequency drops were compared. The crack propagation in the component could then be related to the frequency drop due to existence of the notches. Figure 17 shows the setup for resonant bending fatigue tests where a crankshaft

section is attached to two heavy steel tines and the system acts as a large tuning fork. The right tine is then excited by a shaker. In the second study by Chien et al. (7), as described previously, the effect of residual stress on fatigue analysis of crankshafts was investigated experimentally and numerically. Similar test arrangement and component as in the study of Yu et al. (20) was used, but here the fatigue life of the component was compared for rolled and un-rolled crankshafts. The four-bubble criterion was implemented where cyclic bending loads cause cracks to open and close, and consequently, create oil bubbles as observed in the experiment. Based on the fourbubble failure criterion, crankshafts with the existence of the residual stresses due to the fillet rolling process failed on the order of one million cycles under a bending moment of 508.4 N m (4500 lb-in.). As described earlier, Regulskii et al. (19) performed fatigue tests on a ductile cast iron crankshaft of a motorcycle engine shown in Figure 18 using machines with a crank mechanism for the excitation of alternating stresses as shown in Figure 19. The cantilever mounted crankshaft fragments were tested in bending on a machine with inertial load excitation by rotating unbalanced masses through the connecting rod and crosshead in Figure 19. According to them, the shortcoming of the cantilever mounting of a crankshaft fragment in bending tests is that the loading pattern in such a mounting differs significantly from that in service, although it is in principle applicable for comparative tests. By fragmenting the crankshaft, a number of masses connected to the system are removed affecting the dynamic response of the system. In addition, as applied to the tests of motorcycle engine crankshafts, this arrangement provides no reduction in the number of the specimens tested, since only one specimen can be cut out from a crankshaft. Also, because of its small length, the whole crankpin would be in the grips, in which case, the fracture zone can not be monitored. They developed a special device for mounting the crankshaft in the working area of the machine orthogonally toward the axial force produced by an elastic dynamicdisplacement converter for this force to be applied to the central cheek of the crankshaft halfway between its supports. The main elements of the device itself are two spring plates and a demountable clamp, which encircles the central crank cheek and is connected to the active grip. The elastic converter enables converting the angular displacements of the crosshead with weights into the axial displacements of the active grip of the machine. The elastic converter of the dynamometer connected to the passive grip performs a reverse conversion, i.e., it converts the axial displacements of the passive grip back into the angular displacements of the indicator of the load range. The elastic spring plates, to which the main journals of the crankshaft are fastened, possess a reasonably high stiffness in tensioncompression but rather low bending stiffness, which is by more than an order of magnitude lower than that of the crankshaft. The scheme chosen provides identical loading conditions for both crankpins when the position of the axis along which the load acts remains unchanged. The loading condition enabled obtaining the patterns of in-service fracture of crankshafts, i.e., across the central cheek in laboratory tests. However, it does not provide for bringing the tested crankshaft fragments to separation because redistribution of the bending moments taken up by the crankpins and main journals takes place as the crack grows.

10

The tests were conducted with a fully reversed load cycle with a frequency of 20 Hz at constant load amplitude. The appearance of the first fatigue crack of length from 1.0 to 1.5 mm was taken as the criterion for the number of cycles to failure. The load amplitude of Pa = 24 kN corresponding to the amplitude of the nominal bending stress of a = 88 MPa was chosen so that the mean life of the crankshaft did not move into the low cycle range of lives and that the results enabled plotting a portion of the fatigue curve. Another arrangement for fatigue testing of crankshafts was used by Matsumura et al. (21). In this arrangement, samples were cut from the crankshafts and fixed at both ends, and load was applied at a center journal, as shown in Figure 20. To determine fatigue strength of a V8 engine crankshaft, Jensen (2) outlined and described a crankshaft strength evaluation program as shown in Figure 21. According to this program, the load determination work begins with the sections of the crankshaft to be analyzed. The critical sections of production V8 crankshafts are shown in Figure 22. Small holes are drilled along the axial centerlines of the main journals and crankpins to provide a path for the strain gage lead wires to the front of the crankshaft. Strain gages for bending and torsion are placed at the fillet and crankpin areas, respectively, as shown in Figure 23. This arrangement is necessary to ensure accuracy of measurement due to the specific loading mode. The crankshaft is carefully assembled in the engine and the engine is installed on a dynamometer stand. The gages are calibrated while the engine runs and stops consequently. Separate and combined bending and torsion loads are determined using this arrangement. Following load determination, the crankshaft is cut to allow mounting on the individual test sections in the tuning forks. Due to lower relative importance of torsion compared to bending loads as discussed previously, only bending tests were conducted on crankshafts. Two test pieces containing the critical sections are cut from each crankshaft to be tested. Since experiments show the highest stresses at the critical sections to occur in the crankpin fillets, the main journal fillets in the overlap area were peened to prevent failure. Without peening, failure would occur in the main journal fillets due to the very large stress concentrations introduced by the proximity of the rigid tuning forks. The tuning fork is driven by an electromagnetic vibrator mounted horizontally in a support frame. The vibrator motion is transmitted to the tuning fork through a small drive rod attached to one of the tuning fork arms. The fatigue data related to the study are plotted in the form of a S-N curve in Figure 24. The slope of the line is obtained by running the fatigue samples at several load levels. The measured and experimental data are then compared to obtain the crankshaft factor of safety.

GEOMETRY OPTIMIZATION Mikulec et al. (22) discuss the calculation steps to optimize geometry of a vehicle crankshaft using a powertrain computer model developed at Ford Powertrain and Vehicle Research Laboratory. According to this program, in addition to material

11

properties, a number of design parameters should be specified as input to the optimization problem, as well as to analyze crankshaft stresses, torsional frequency and fatigue safety factor. These include crankshaft reciprocating weight, rotating weight, main cheek width, main cheek thickness, bore spacing, connecting rod length/crank radius, number of cylinders, arrangement of cylinders, number of main journals, number of crank-pins, main fillet radius, rolled fillet fatigue strength improvement (due to rolling), and cheek relief thickness. Diameter of main journals, length of main journals, diameter of crank-pins and length of crank pins are considered as design variables. The constraints of the optimization problem are named as crank-pin pressure due to peak cylinder pressure or high speed inertia load, fatigue safety factor of the main cheek, first mode torsional natural frequency, and the distance from oil hole to the nearest fillet. Henry et al. (16) modified the web design of a nodular cast iron diesel engine crankshaft based on the crankshaft durability assessment procedure they implemented, as described previously (i.e. Figure 11), to reduce maximum stress within the bounds of reasonable geometries. This constraint allowed changes in the web width and pin side web profile only. The initial and modified webs are shown in Figure 25. This modified design gave improvements in fatigue safety factor by 19% for the crank pin fillet studied. They attributed the stress reduction to two mechanisms, the first to be the geometric reinforcement of the webs, and the second to be the decrease in the maximum main bearing reaction on the fifth journal due to the induced change in crank balancing.

COST ANALYSIS A systematic cost estimation of crankshafts is provided in the work of Nallicheri et al. (23). Dividing the cost of crankshafts into variable and fixed cost, they evaluate and compare the production cost of crankshafts made of nodular cast iron, austempered ductile iron, forged steel, and microalloyed forged steel. The common variable cost elements are named as the costs of material, direct labor, and energy. The common elements of fixed cost are named as the costs of main machine, auxiliary equipment, tooling, building, overhead labor, and maintenance. Based on a cost model and assuming estimation parameters for a particular crankshaft made of the four materials and manufacturing processes, the production cost was obtained as shown in Figure 26 and Table 6 for a production rate of 792,000 parts per year. While the material costs of all the four processes are essentially similar, the labor cost contribution in the steel forging case is higher than that of nodular castings due to the more complex machining process for steel. The breakdown shows that the ADI and forged steel crankshaft costs are similar, with steel forging being 7% lower in cost than ADI. Sensitivity of the cost with respect to volume is evaluated and shown in Figures 27 and 28. Since the use of microalloyed steel eliminates the need for heat treatment, it yields savings compared with other processes. Moreover, using microalloyed steel reduces machining costs compared to conventional forging due to eliminating part or all of heat treatment steps. Nallicheri et al. (23) conclude that when designers are looking for high strength crankshafts, microalloyed forging steels are cost effective, high performance

12

replacements for nodular castings. ADI crankshafts were considered cost effective in low production runs (below 180,000 parts per year). According to them, the choice of a cost effective production route for crankshafts is dependent upon two factors; the production volume and the requirements of the engine. If the engine design can accommodate the properties offered by a nodular cast crankshaft, then the cast crankshaft offers the most cost effective fabrication route. If the design calls for more stringent strength requirements, then the other alternatives must be considered. At production volumes above 200,000 parts per year, microalloyed steel forgings offer the most cost effective, high performance crankshafts. It should be noted that more emphasis on fuel efficiency and the rapidly growing need for optimization in recent years, places stringent performance requirements on engine components, particularly the rotating parts such as crankshafts. Therefore, microalloyed steel forgings offer the best potential for achieving fuel efficient and high performance engines. Hoffmann and Turonek (11) examined the cost reduction opportunities for forged steel crankshafts, with raw material cost, fatigue strength and machinability being the primary factors evaluated. In their study, a cost model, supported by fatigue and machinability data, was utilized to select the lowest cost among the material grades they examined including carbon steel (CS), alloy steel (AS), high strength steel (HS), and microalloyed steels (MA1 and MA2), as previously discussed and listed in Table 1. Table 7 shows a typical cost breakdown for the carbon steel crankshaft computed by the previously developed cost model. The principal assumptions to determine the cost were steel cost (CS=$0.517/kg, CS-HS=$0.546/kg, AS=$0.572/kg, AS-HS=$0.600/kg, and MA1/MA2=$0.539/kg), machining waste (35% for four cylinder and 15% for six cylinder), and the finished weight (16.8 kg for four cylinder and 21.4 kg for six cylinder). Tables 8 and 9 illustrate costs at three production levels for both the four cylinder and six cylinder crankshafts. Table 8 shows that replacing carbon steel with high strength carbon steel has no cost benefit. The savings realized by using better machining steel are offset by the higher cost of the raw material. Table 9, however, shows that for the higher strength application requiring alloy steel there are savings to be gained by using high strength alloy steel. Both tables show that maximum savings are realized from the microalloy grades. These savings are summarized in Table 10. The microalloy grade could reduce the finished cost by 11 to 19 percent compared to the quenched and tempered alloy steel (SAE 4140), and by 7 to 11 percent compared to the quenched and tempered carbon steel (SAE 1050) of their study. These microalloy grades meet or exceed the fatigue strength of the original materials for the application studied, and have better machinability characteristics. Table 11 shows the potential cost reduction that could be obtained both in machining and raw material (4.9 kg saved) by redesigning the four-cylinder crankshaft with cheeking (machining the sides of the counterweights) and topping (machining the ends of the counterweights) eliminated. These savings are 7 percent for a carbon steel design and 13 percent if the material is changed to a microalloy grade. In their study, high sulfur levels (0.1%) for quenched and tempered medium carbon and alloy steels, that enhance machining, show no appreciable reduction in fatigue strength. For alloy steels this has a potential cost saving of 2-6 percent, while for carbon steel the study indicated no savings.

13

CONCLUSIONS 1. Crankshaft is one of the most critically loaded components of internal combustion engine and experiences cyclic bending and torsion loads during its service life. The main sources of loading are gas load due to the combustion process transmitted to the crankshaft by connecting rods, and inertia load due to the mass of the component and its attachments. 2. Fillets in the crankshaft act as stress raisers and endure the highest level of stress under service loading. They are the critical locations, where cracks can nucleate at a fillet surface due to combined cyclic bending and torsion. 3. Failure sources of crankshafts include oil absence, defective lubrication on journals, high operating oil temperature, misalignments, improper journal bearings or improper clearance between journals and bearings, vibration, high stress concentrations, improper grinding, high surface roughness, and straightening operations. 4. Crankshafts are typically manufactured from forged steel, nodular cast iron and austempered ductile iron (ADI). When forged steels are compared to cast iron and alloyed ductile iron used in crankshafts, the fatigue properties of forged steels are generally found to be better that that of cast iron. 5. Manufacturing by forging has the advantage of obtaining a homogeneous part that exhibits less number of microstructural voids and defects compared to casting. In addition, directional properties resulting from the forging process helps in higher strength in the grain-flow direction. Grain flow pattern is an indicator of forging quality as well as directional strength. 6. Fillet rolling has traditionally been used to induce compressive residual stresses at the crankshaft fillets, making the process of fillet rolling quite beneficial to fatigue strength. Using optimum conditions for fillet rolling will improve the fatigue strength of the filleted specimens equal to or even above that of the smooth specimens. 7. Shot peening also improves the fatigue strength of the fillet, but to a lesser degree than fillet rolling. Although the compressive residual surface stress by shot peening can be higher than that of fillet rolling, the depth of penetration of the compressive stress developed by shot peening may be shallower than by fillet rolling. 8. Induction surface hardening is also used to induce compressive residual stresses at the critical locations of the crankshaft. A challenge in induction surface hardening is to ensure a very slight variation in hardness and the existence of compressive residual stresses with proper magnitude and distribution in transition areas to the hardness of the base material. 9. Because of the high cost of the component and test systems and complex geometry of multiple-bearing one-piece crankshafts, fractions of the crankshafts consisting of one crank with two coaxially located main journals fixed in the grips of a testing machine are commonly used for fatigue testing. Due to lower relative importance of

14

torsion compared to bending loads, only bending tests are often conducted on crankshafts. 10. Diameter of main journals, length of main journals, diameter of crank-pins and length of crank pins have been considered as design variables in optimizing the geometry of crankshafts. Crank-pin pressure, fatigue safety factor of the main cheek, first mode torsional natural frequency, and the distance from oil hole to the nearest fillet have been used as the constraints in such optimization. 11. The use of microalloyed steel eliminates the need for heat treatment, resulting in cost savings compared to conventional forging of quenched and tempered alloy or carbon steels, while meeting or exceeding the fatigue strength of these steels. In addition, elimination of the heat treatment reduces machining costs. 12. The choice of a cost effective production route for crankshafts is dependent upon the production volume and the requirements of the engine. For high performance crankshafts with stringent strength requirements and at high volumes, microalloyed forging steels are most cost effective. ADI crankshafts may be cost effective in low production runs. Cast crankshafts can be cost effective in high volume productions, only if the engine design can accommodate the lower strength.

REFERENCES 1. Shamasunder, S., 2004, Prediction of Defects and Analysis of Grain Flow in Crank Shaft Forging by Process Modeling, SAE Technical Paper No. 2004-01-1499, Society of Automotive Engineers. 2. Jensen, E. J., 1970, Crankshaft Strength Through Laboratory Testing, SAE Technical Paper No. 700526, Society of Automotive Engineers. 3. Silva, F. S., 2003, An Investigation into the Mechanism of a Crankshaft Failure, Key Engineering Materials, Vols. 245-246, pp. 351-358. 4. Dubensky, R. G., 2002, Crankshaft Concept Design Flowchart for Product Optimization, SAE Technical Paper No. 2002-01-0770, Society of Automotive Engineers. 5. Grum, J., 2003, Analysis of Residual Stresses in Main Crankshaft Bearings after Induction Surface Hardening and Finish Grinding, Journal of Automobile Engineering, Vol. 217, pp. 173-182. 6. Ren, W., Keyu. L., and Lee, Y. L., 2004, Optical Measurement of Residual Stress at the Deep-Rolled Crankshaft Fillet, SAE Technical Paper No. 2004-01-1500, Society of Automotive Engineers. 7. Chien, W. Y., Pan, J., Close, D., and Ho, S., 2005, Fatigue Analysis of Crankshaft Sections Under Bending with Consideration of Residual Stresses, International Journal of Fatigue, Vol. 27, pp. 1-19.

15

8. Park, H., Ko, Y. S., and Jung, S. C., 2001, Fatigue Life Analysis of Crankshaft at Various Surface Treatments, SAE Technical Paper No. 2001-01-3374, Society of Automotive Engineers. 9. Gligorijevic, R., Jevtic, J., Vidanovic, G., and Radojevic, N., 2001, Fatigue Strength of Nodular Iron Crankshafts, SAE Technical Paper No. 2001-01-3412, Society of Automotive Engineers. 10. Gundlach, R. B., Semchyshen, M., and Whelan, E. P., 1998, Notch Sensitivity and Fatigue in Austempered Ductile Iron, SAE Technical Paper No. 980685, Society of Automotive Engineers. 11. Hoffmann, J. H. and Turonek, R. J., 1992, High Performance Forged Steel Crankshafts - Cost Reduction Opportunities, SAE Technical Paper No. 920784, Society of Automotive Engineers. 12. Chatterley, T. C. and Murell, P., 1998, ADI Crankshafts An Appraisal of Their Production Potentials, SAE Technical Paper No. 980686, Society of Automotive Engineers. 13. Pichard, C., Tomme, C., and Rezel, D., 1993, Alternative Materials for the Manufacture of Automobile Components: Example of Industrial Development of a Microalloyed Engineering Steel for the Production of Forged Crankshafts, In Proceedings of the 26th ISATA International Symposium on Automotive Technology and Automation, Aachen, Germany. 14. Payer, E., Kainz, A., and Fiedler, G. A., 1995, Fatigue Analysis of Crankshafts Using Nonlinear Transient Simulation Techniques, SAE Technical Paper No. 950709, Society of Automotive Engineers. 15. Prakash, V., Aprameyan, K., and Shrinivasa, U., 1998, An FEM Based Approach to Crankshaft Dynamics and Life Estimation, SAE Technical Paper No. 980565, Society of Automotive Engineers. 16. Henry, J., Topolsky, J., and Abramczuk, M., 1992, Crankshaft Durability Prediction A New 3-D Approach, SAE Technical Paper No. 920087, Society of Automotive Engineers. 17. Guagliano, M. and Vergani, L., 1994, A Simplified Approach to Crack Growth Prediction in a Crankshaft, Fatigue and Fracture of Engineering Materials and Structures, Vol. 17, No. 11, pp. 1295-1994. 18. Guagliano, M., Terranova, A., and Vergani, L., 1993, Theoretical and Experimental Study of the Stress Concentration Factor in Diesel Engine Crankshafts, Journal of Mechanical Design, Vol. 115, pp. 47-52. 19. Regulskii, M. N., Pogrebnyak, A. D., and Balakovskii, O. B., 2002, Procedure and Results of Investigation into Fatigue Strength Characteristics of Motorcycle Engine Crankshafts, Strength of Materials, Vol. 34, No. 6, pp. 629-635. 20. Yu, V., Chien, W. Y., Choi, K.S., Pan, J., and Close, D., 2004, Testing and Modeling of Frequency Drops in Resonant Bending Fatigue Tests of Notched Crankshaft

16

Sections, SAE Technical Paper No. 2004-01-1501, Society of Automotive Engineers. 21. Matsumura, Y., Kurebayashi, Y., Konagaya, D., and Mizuno, K., 1999, Development of Nitrocarburizing Steels for Crankshafts, SAE Technical Paper No. 1999-01-0601, Society of Automotive Engineers. 22. Mikulec, A., Reams, L., Chottiner, J., Page, R. W., and Lee, S., 1998, Crankshaft Component Conceptual Design and Weight Optimization, SAE Technical Paper No. 980566, Society of Automotive Engineers. 23. Nallicheri, N. V., Clark, J. P., and Field, F. R., 1991, Material Alternatives for the Automotive Crankshaft; A Competitive Assessment Based on Manufacturing Economics, SAE Technical Paper No. 910139, Society of Automotive Engineers. 24. Halderman J. D. and Mitchell, C. D., 2001, Automotive Engines, 1st edition, Prentice Hall, Inc. 25. Payer, 2000, Advanced Numerical Simulation Techniques for the Fatigue and NVH Optimization of Engines, 2nd MSC Worldwide Automotive Conference.

17

Table 1

General steel grades tested, heat treatments and hardness in the study by Hoffmann and Turonek (11).

Table 2

Hardness and bending fatigue strengths for the steels studied by Hoffmann and Turonek (11).

Table 3

Mechanical properties from crankshaft journals studied by Hoffmann and Turonek (11).

18

Table 4

Summary of fatigue test results on forged and ADI crankshafts studied by Chatterley and Murrell (12).

Table 5

Fatigue experiment results on specimens from competing crankshaft materials studied by Pichard et al. (13).

19

Table 6

Breakdown of cost by manufacturing factor for fully machined crankshafts of alternative processes in 1991 currency from the study by Nallicheri et al. (23).

Table 7 Cost breakdown for four cylinder carbon steel crankshaft based on annual volume of 100,000 from the study by Hoffmann and Turonek (11).

20

Table 8

Cost data for the four cylinder crankshaft of the study by Hoffmann and Turonek (11).

Table 9

Cost data for six cylinder crankshaft of the study by Hoffmann and Turonek (11).

21

Table 10 Summary of potential savings for four and six cylinder engines of the study by Hoffmann and Turonek (11).

Table 11 Cost reduction by eliminating cheeking and topping for the four cylinder crankshaft of the study by Hoffmann and Turonek (11).

22

Figure 1

Typical crankshaft with main journals that support the crankshaft in the block. Rod journals are offset from the crankshaft centerline (24).

Figure 2

An automotive ductile cast iron crankshaft and a close-up of the fillet (7).

23

Figure 3

Load data for the critical section (one of the fillets) of a V8 engine nodular iron crankshaft versus crank angle (2).

Figure 4 Crankshaft design tasks presented by Dubensky (4).

24

(a)

(b) Figure 5 (a) Casting and (b) forging process flows for manufacturing crankshafts (23).

Figure 6

Mechanical machining and heat treatment procedure from blank to crankshaft (5).

25

Figure 7

S-N curve rig test results with various surface treatments for CrMo alloy steel (8).

26

Figure 8

Ductile iron crankshaft casting used for smooth and notched test specimens in the study by Gundlach et al. (10).

Figure 9

Comparison of fatigue strength of the original, fillet-rolled, shot-peened, and smooth wall ground ADI specimens in the study by Gundlach et al. (10).

27

Figure 10 Stress distribution due to maximum gas load at the fillet of a crankshaft (25).

Figure 11 Crankshaft durability assessment procedure used by Henry et al. (16).

28

Figure 12 Solid model of the diesel engine crankshaft in the study of Henry et al. (16).

Figure 13 The calculated operating conditions in the study of Henry et al. (16).

Figure 14

Stress cycle for the crank pin fillet in the study of Henry et al. (16).

29

Figure 15 Numerical versus experimental axial strains and pattern along the crank fillet (axis unit in mm) obtained by a plane model (18).

Figure 16 Results of constant amplitude fatigue testing of crankshafts: (1) nonhardened crankshafts without a fillet; (2) crankshafts without a fillet subjected to vibration surface hardening; (3) non-hardened crankshafts with a fillet (19).

30

Figure 17

Experimental setup for resonant bending fatigue tests of crankshaft (20, 7).

31

Figure 18 Schematic diagram of a crankshaft of a motorcycle engine in the study by Regulskii et al. (19).

Figure 19 Scheme of mounting a full scale motorcycle crankshaft on elastic elements simulating two crankshaft bearings (a) and block diagram of a test bench for testing crankshafts in bending in the crank plane (b) in the Regulskii et al. study: (1) elastic converter of torsional vibrations into axial ones; (2) active grip; (3) fragment of the crankshaft; (4) passive grip; (5) elastic dynamometer; (6) optic monitor; (7), (8), and (10) diaphragms; (9) spring plate; (11) crosshead; (12) connecting rod of the dynamic displacement exciter (19).

32

Figure 20 Schematic view of bending fatigue test of a crankshaft (21).

Figure 21 Crankshaft strength evaluation program by Jensen (2).

Figure 22 V8 crankshaft critical sections in the study by Jensen (2).

33

Figure 23 Strain gage bridge placement in the study by Jensen (2).

Figure 24 S-N diagram of crankshaft section strengths in the study by Jensen (2).

34

Figure 25 Initial (left) and modified web designs to reduce maximum stress within the bounds of reasonable geometries in the study by Henry et al. (16) this constraint allowed changes in web width and pin side web profile only.

Figure 26 Cost breakdown for crankshafts of alternative processes based on 792,000 units per year in 1991 currency in the study by Nallicheri et al. (23).

35

Figure 27 As-formed piece costs as a function of annual production volume in the study by Nallicheri et al. (23).

Figure 28 Finished piece costs alternative processes as a function of annual production volume in the study by Nallicheri et al. (23).

36

Potrebbero piacerti anche