Stamping and Form Ability

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 256

Stamping And Formability: Automotive Sheet Metal Stamping And Formability (January 1989)

PREFACE
This state-of-the-art report on forming sheet metals for automotive applications is one in a series of reports on technical subjects of common interest to steel suppliers and automotive users. Other topics covered in this series include fatigue, corrosion fatigue, and welding. Two major purposes are served. First, as a thorough compilation of knowledge on forming sheet metals, it assembles information in a convenient source document for use by the researcher and forming specialist. As such, it also provides source information for a text book for the practitioner and for briefer summaries in specific areas for beginner. Second, areas of needed research are identified by an analysis of gaps in our understanding of sheet metal formability. Comments by users of this report will be welcome. AISI is grateful to Stuart Keeler, formerly with National Steel Corporation and now with the Budd Company, for his intensive study; to Stephen Denner of National Steel for his valuable cooperation; to Nancy DeSmet of National Steel for typography, and to Bill Reyman for artwork and publication design.
AMERICAN IRON AND STEEL INSTITUTE Automotive Applications Committee

This publication is for general information only. The information in it should not be used without first securing competent advice with respect to its suitability for any given application. The publication of the information is not intended as a representative or warranty on the part of the American Iron and Steel Institute or any other person named herein that the information is suitable for any general or particular use or of freedom from infringement of any patent or patents. Anyone making use of the information assumes all liability arising from such use. Produced by W.P. Reyman New York Copyright American Iron and Steel Institute 1990 This report reviews the State-of-Art of Automotive Sheet Metal Formability. It was commissioned by the American Iron and Steel Institute, whose support is gratefully acknowledged. The purpose of this project was to obtain and study the available knowledge base of automotive sheet steel formability - both published and unpublished. By limiting the scope of the study to automotive sheet steel, the types of steel, the thicknesses, and the shapes included in this study were likewise limited. The material reviewed is in this report, which becomes both a textbook in the field of sheet metal formability and a catalog of available information. As a textbook, this document attempts to provide the reader with an overview of the available knowledge. As an overview, the report does not provide all of the necessary background information, but assumes a minimum level of knowledge and understanding of the fundamentals of sheet metal formability. the reader, therefore, is required either: 1) to have the basic knowledge of the field of sheet metal formability or 2) to obtain and study supplementary references for basic information in the field. This document will be used most often, however, as a handy catalog of available information, especially key figures and references. These can be readily found within this single document without scanning a number of different papers and books.

This report is organized such that individual sections (bolded in the Table of Contents) can be separated from the report and utilized individually with a minimum of additional effort. To this end, the pages and figures of each section are numbered independently. In some cases, key figures are duplicated in more than one section to facilitate this separation. This report is organized such that individual sections (bolded in the Table of Contents) can be separated from the report and utilized individually with a minimum of additional effort. To this end, the pages and figures of each section are numbered independently. In some cases, key figures are duplicated in more than one section to facilitate this separation. The author would like to thank the staff of National Steel Product Application Center for their constructive reviews of each section and for their numerous suggestions. Special appreciation goes to Dr. Stephen Denner for this guidance and motivation which made completion of this extensive project possible. Sheet metal forming is an interactive system. For many years, this important fact has received tacit acknowledgment. However, only within the last decade has this fact been truly incorporated into our understanding of sheet metal formability and used to exploit the positive features of this interaction to improve sheet metal formability. Two general descriptions of the forming system can be found. While similar, each defines the system from a different perspective. The first, and most common, defines the system in terms of the readily identifiable components which make up the forming system (K-23). These components, shown in Figure 1-1 are: DIE DESIGN MATERIAL LUBRICANT PRESS Sometimes five components are listed; part design is added to the list. While part design and die design are closely interrelated, the two sometimes are separated for the purpose of studying the interactions of the forming system. For example, each die designer may have a different approach to forming a given part and each may design a different die to form the part. One die designer may attempt to create the part geometry by pure stretch forming, while another may attempt to create most of the required length of line by pulling metal from the binder. Thus, one important aspect of forming system analysis is to determine whether the initial part design exceeds all possible forming limits or whether the die design is incorrect for the specific part. Some die designs permit a difficult part to be easily made. Other die designs create a difficult problem out of an easy part design. The other components of the forming system are readily appreciated. They represent the obvious factors which affect the ability of the sheet of metal to be transformed into the specified geometry. This is one basis for defining and understanding sheet metal formability. In this review, it is assumed that a given sheet of metal does not have an absolute level of formability which can be measured and defined. Instead, sheet metal has certain properties or parameters which can be measured and which can be related to formability requirements within the context of the specific system. This concept can best be understood by several examples. Consider, for example, sheet steel thickness. A common rule of thumb is that the formability of steel increases with increasing thickness. When this steel is inserted into a pure stretch forming operation, a thicker steel will, in fact, product a higher dome. However, when this steel is inserted into a draw operation, the increased thickness can cause insufficient clearance between the punch and die, as well as cause a jam in a draw bead. The center of the stamping would be ripped out very early in the stroke of the punch. Having performed these two experiments, should it be concluded that sheet steel formability increased or decreased with increasing sheet thickness?

In another example, a given sample of steel in a given die with a given lubricant will break when the combination is placed in a press line A but will combination is placed in press line D. What absolute level of formability can be assigned to the sheet steel? Finally, a third example is the normal anisotropy ration, rm, of the steel (See Section 4.1.4.8). Cup drawing improves with an increase in the normal anisotropy ration (A-28), while pure stretching over a hemispherical punch decreases with an increase in the normal anisotropy ratio (H-21). Therefore, formability of sheet steel relative to normal anisotropy can not be simple defined. Instead, the formability of sheet steel can be defined only relative to the mode of deformation and in terms of all other boundary conditions imposed by the entire forming system. Thus, the definition of the forming system and the breakdown of a complex stamping into component forming modes (Section 2) becomes an important keystone in the understanding of sheet metal formability. Likewise, identification of the major sheet metal forming process variables is important. Siekirk recently identified thirty such variables (S-24) and the list is constantly growing as the forming system is studied further. In the same manner, the ability of a single simulative test to prescribe a single formability rating has been challenged (L-10). A somewhat related description of the forming system (B-2) is shown in Figure 1-2. Here the interactive components are: INCOMING METAL TOOL-METAL INTERFACE DEFORMATION ZONE COMPLETED PART This description of the generalized forming system is commonly used in universities and can be applied, not only to sheet metal forming, but to bulk metal forming, such as extrusion, forging, swagging, and slab rolling. The scope of this forming system analysis generally is rather narrow and focused. The emphasis here is not the shape of the exiting stamping and its proximity to failure but the properties of the exiting metal. Analyses are made to determine the modification of the incoming properties resulting from the deformation process to predict the strength of the completed part. Both of the previous descriptions of the forming system assume that a full understanding of the system, its components, and all the interactions is possible. Such is the foundation of the mathematical simulations of the forming system currently being formulated (See Section 8.2). Knowing how the system works allows a prediction of the final output of the system (the completed stamping) without even constructing the system. Unfortunately, a more realistic description of the system (K-23) is provided in Figure 1.3. The interactive forming system is so complex that accurate description or understanding of the system is not possible. Thus, black box analysis techniques are necessary. Here the total forming system with its interactions is unknown; this is indicated simply by a black box which represents the total sum of the actions and interactions. All that is known and can be measured is the input and output conditions of the metal. In this analysis procedure, the forming system has a series of dials on the side of the black box. The goal is to select one dial which will create the desired output conditions. Only one dial is changed at a time. The first dial is turned to the left. If the stamping is improved, the new setting is used. If the stamping is worse, the first dial is turned to the right. If the stamping is improved, the new setting is used. If the stamping is worse, the first dial is reset to its initial position and the second dial is tried. The dials can be any factors within the system which may affect the output of the system - lubrication, blank size, holddown load, die radius, ram position, sheet surface roughness, work hardening exponent (n), or any other factor. The secret of success is selecting the correct dial or dials. The inexperienced experimenter may have to try 125 dials before the correct one is found. The experienced tool and die setter may have to try 15 dials. On the other hand, either of them could select the correct dial the first time by chance.

Such is the current trial-and-error procedures commonly used in the press shop. Best guesses are made, based on past experience with similar stampings, as to what modifications are required to produce successful stampings. The problem becomes more complex, however, with the introduction of new steels (electrogalvanized and high strength steels), new lubricants (demand lubricants which change frictional conditions in response to localized heat and pressure), and new forming techniques (drawing into high pressure tanks) for which no experience base has been accumulated. Fortunately, a system of measuring the system response is available. This is the circle grid analysis system (Section 8.1.1). The effect of all system modifications can be monitored without breakage occurring. In addition, Forming Limit Diagrams (Section 4.2) can be used to assess the severity of any combination of forming system parameters and monitor the direction of any changes. Acknowledgment of the forming system is critical to the understanding of sheet metal formability. This state of the art review, therefore, is organized according the main components of the forming system:
PART/DIE DESIGN Section 3

MATERIAL LUBRICANT PRESS


SYSTEM INTERACTIONS

Section 4 Section 5 Section 6


Section 8

The theme of each section is how that section interacts with the other components of the forming system. Figure 1-1: One description of the interactive forming system includes those components readily identifiable within the press shop (K-23).

Figure 1-2: Generalized deformation processing system represented by a roll forming operation (B-2). The four zones are incoming metal, tool-metal interface, deformation zone, and completed

part.

Figure 1-3: Interactive forming system represented by an input-output black box is commonly used to describe the current trial-and-error modification of forming processes (K-23).

2.1 Introduction
Forming operations convert coils of steel received from the steel mill into stampings. Rarely used in their as-formed condition, these stampings usually are subsequently assembled with other stampings, or parts, by welding, bonding, or mechanical fastening. These subsequent operations are important in that they place constraints on the forming operations. For example, welding may require a weld flange of minimum width which must be buckle free.

In addition, attempts are being made to combine more and more individual stampings into one single stamping to compete with the design trends being established by the plastics industry. Finally, reduction in sheet metal thickness for weight reduction and cost considerations has required the addition of embossments, ribs, and other design features to maintain part stiffness and other performance requirements. All these factors have greatly increased the complexity of the average stamping. To study the formability of sheet metal as it relates to these complex stampings, it becomes necessary to break down the complex stamping into its component sections (E-1, E-2, A-12, K-26, A-10, M-10). Two methods are used to divide a complex stamping into its component sections. The first method is by the geometry of the part; this method generally is used by the stylist and part designer. Here the final geometry and dimensions of the functional unit are described independent of how the geometry is to be generated. This geometry usually describes the final panel or part generated by the stamping operations prior to assembly. The second method is to subdivide the stamping into components by forming operations or forming mode; this technique is used by the die designer to generate the required geometry. It consists of an initial stamping which may be preceded by one or more preforms and the followed by a number of additional restrike, trim, flange, punch, and other operations before it becomes a finished part. The initial stamping may or may not resemble the final part. In some extremes, more than fifty percent of the initial blank is removed as offal. A complex part representative of geometries commonly found in typical automotive body panels is sketched in Figure 2-1. The geometry of the part can be described by specifying the dimensions of the top surface, the side walls, the corners, and the flanges. Added to this overall stamping geometry are sub areas, such as embossments, holes, slots and other functional zones. The part designer thinks in terms of the required geometry to accomplish the required function or the geometry needed to fulfill the styling shape usually without concern about how the part is to be made. Definition of this geometry is not difficult. For example, the part print for Figure 2-1 could require the lower left corner radius to be 1 times metal thickness (1t) or 15 times metal thickness (15t). The 1t radius may actually be required for clearance or other purposes or it may be simply an arbitrary number put on the part print because it looks crisp. The 1t radius, however, restricts the allowable depth of the initial stamping and probably will require two or three stamping operations to generate the specified part surface. A 1t radius corner may even be impossible to produce or be cost prohibitive in some part configurations. The second method of breaking down a complex part is by its forming operations or forming modes (Figure 2-2). Typical forming operations are defined in the remainder of this section. A discussion on design details is included in Section 3. It is important to note that specific geometrical shapes can be created by more than one forming operation or mode. This can be important to circumvent some forming limits. For example, an embossment followed by one or more redraws can be substituted for an impossible stretch operation.

2.2 Cutting
The preparation of the sheet metal blank or workpiece from coils, strips, or sheets by a cutting operation is an easily identified first step in the formation of almost all sheet metal stampings. In a few cases, the blank itself is the final completed part, such as a motor lamination. In most cases, though, the blank is subjected to subsequent forming. Cutting operations is one plane are classified by the terms shearing, blanking, slitting, piercing, and lancing. Cutting operation in more than one plane are classified by the terms trimming and parting.

SHEARING is performed by a blade acting along a straight line. The work metal is placed between a stationary lower blade and a movable upper blade and is severed by bringing the blades together. Nondeveloped blanks are generated by shearing. BLANKING involves a cutting action about a closed shape which is the piece retained for further processing. The closed shape may be composed of any number of straight and curved line segments. Developed or contoured blanks are generated by blanking. SLITTING is the cutting of lengths (usually coils) of sheet metal into narrower lengths by means of one or more pairs of circular knives. This operation often precedes shearing or blanking and is used to produce exact blank or nesting widths. PIERCING is forming a hole in sheet metal with a pointed punch with no metal fallout. LANCING makes an opening without completely separating the cut piece from the body of the metal sheet, such as for louvers. TRIMMING removes unwanted metal from the finished part that was required for some previous stamping operation, such as binder areas, or was generated by a previous stamping operation, such as the earing zone on the top of a deep drawn cup. PARTING operations are used to separate two identical or mirror image stampings that were formed together (typically for the expediency of making two parts at one time or to balance the draw operation of a nonsymmetrical part). Parting also is an operation that involves two cutoff operations to produce contoured blanks from strip. Scrap is produced in the parting operation.

2.3 First Form Operations


First forming operations can be organized and catalogued in a variety of ways. One logical method is based on increasing complexity. LINE SHAPE
Straight Bend Bend-and-Straighten Plane Strain Curved Shrink Flange Stretch Flange Hole Expansion

SURFACE SHAPE
Infinite Blank Embossment Biaxial Stretch Limited Blank Draw

Each of these forming operations will be described in detail.

2.3.1 BEND
Bending is one of the most common methods used to change the shape of sheet metal. Two types of bends are commonly employed (Figure 2-3). The first is know by several names, such as the free bend, v-bend, U-bend, and press bake bend. In each case a punch forces the metal into a long channel die as both free edges swing upward. In the second type of bend, known as a wiping bend, one edge is held securely while the punch wipes or swings the free edge down.

Bending is one of the most common methods used to change the shape of sheet metal. Two types of bends are commonly employed (Figure 2-3). The first is know by several names, such as the free bend, v-bend, U-bend, and press bake bend. In each case a punch forces the metal into a long channel die as both free edges swing upward. In the second type of bend, known as a wiping bend, one edge is held securely while the punch wipes or swings the free edge down. All bending operations have a number of common characteristics. (a) One or more metal edges swing through space. This swing must be calculated to allow for sufficient room in the tooling. (b) The metal bends along a narrow line that acts much like a plastic hinge. This line is the axis of the bend. (c) An element of metal, and corresponding circle grid on the outer surface of the flat blank, would deform in the following manner. The material would be elongated across the bend to generate the required increase in the length line. This elongation would be the largest strain in the surface of the blank. Thus, the major axis of the resulting ellipse would be oriented across the bend and would be longer than the original circle diameter. The strain value is calculated from the formula:
% Strain = If - Ii Ii x 100 I Ii

where li is the initial gage length (initial circle diameter) and lf s the final gage length (axial length of the ellipse). The major axis of the ellipse is longer than the initial circle diameter and creates a positive values of strain, e1, which indicates a tensile strain or elongation. No deformation occurs along the axis of the bend. The minor strain (minimum strain or strain perpendicular to the major strain), e1, therefore is zero because the constant length of the axis does not change the original circle diameter. The bending operation is given the notation (+ by 0) or tensile major strain by zero minor strain. (d) The inner (concave) surface is placed in an opposite deformation state or in compression (Figure 23b). Somewhere between the outer and inner surfaces a neutral axis or a line of zero strain exists. Thus, the bending action creates an extreme strain gradient from the convex to the concave surface. (e) Metal outside the plastic hinge is unstrained and remains in the as-received condition.

2.3.2 BEND-AND-STRAIGHTEN
The final shape generated by a bend-and-straighten operation is identical to that of a bending operation. The intermediate steps are quite different and therefore generate different characteristics in the final product. In the bend-and-straighten operation, the swing of the metal is prevented by a blank holder (Figure 2-4a). At the beginning of the forming operation, metal is wrapped around the bottom radius of the punch by a bending operation. At the same time, metal also is wrapped around the die radius, again by bending. Once the initial configurations of these radii are obtained, the primary forming process can begin. Thereafter, each additional element in the final wall begins initially in the flange and is pulled toward the radius zone. Upon entering the die radius zone (Figure 2-4b), the element is bent to conform to the contours of the radius. When leaving the radius zone, the element must be unbent or straightened to conform to the straight wall. Thus, the operation is accurately called bend-and-straighten. The primary difference between bend and the bend-and-straighten operation is the condition of the final wall. In the bend operations, the final part wall is swung into position and remains in the unworked state. In the bend-and-straighten operation the outer (convex) surface is first placed in tension and then in compression. The inner surface undergoes the reverse sequence. Even though the final wall is straight, the metal has undergone a deformation sequence. Depending on the die radius of curvature, the working sequence can be quite severe. Thus, the properties of the wall are changed.

The circular grid pattern would show the tensile major strain by zero minor strain pattern around the radius to be identical to the bending operation. Little, if any, resultant strain would be detected in the wall.

2.3.3 PLANE STRAIN STRETCH


The plane strain stretch operation is similar to a bending operation but a tensile strain component is added across the radius (Figure 2-5). If the bend radius is large compared to the sheet thickness, the bending strain component can be small compared to a large tensile strain component. This produces a strain condition in which both the outer and inner surfaces show a tensile major strain (though different) by a zero minor strain. This deformation state is commonly found when the ratio of inner bend radius to sheet thickness (r/t) is greater than 10. In contrast, the sheet is considered a membrane with the throughthickness stress being negligible compared to the in-sheet stresses when the ratio r/t exceeds 20. Eary (E-1) describes this operation as stretching a large radius, partial cylinder. This particular type of deformation is very sensitive to the type of die construction. Figure 2-6 shows approximately the same angle being generated in two ways. Depending on the tip or angle of the die, the deformation can be accomplished by plane strain stretch (2-6a) or must be made by a bend-andstraighten operation (2-6b).

2.3.3 FLANGING:
SHRINK AND STRETCH
Another degree of complexity is added to the bending operation by changing the line of bending from a straight line to a curved line. Deformation along the line of bending is not longer zero but may be either tensile (positive) or compressive (negative). One form of flanging is called shrink flanging (Figure 2-7a). As the name implies, the length of the flange shrinks as it is formed. Each radial zone (shaded region) is folded 90 degrees along a radical line to form the flange or wall. Since the arc length A of the final flange or wall is smaller than the arc length B of the element from which it was formed, a compression must take place in the circumferential direction indicated by the arrows. The greater flange the flange depth, the greater the amount of compression required. The compressive strain state is indicated by the minor axis of the ellipse being less than the diameter of the original gridded circle. This compressive strain state has a tendency to produce the buckles or wrinkles common in shrink flanging. A deeper flange has more tendency to buckle because of the greater compressive strain required. The opposite form of flanging is stretch flanging (Figure 2-7b). Here the material is stretched as it is flanged. The initial arc length is shorter than the required arc length of the flange. A tensile strain is needed to generate this required increase in length of line. This elongation requires the major axis of the ellipse to reside in the circumferential direction (a 90 degree rotation from that found in shrink flanging). A greater flange depth requires a greater change in arc length or elongation. The minor strain is negative near the edge (free to pull in as for a tensile test) and zero near the bend line. Since the deformation state is now tensile, no buckling occurs. However, the limit of the operation becomes the tensile deformation limit of the material. Exceeding this limit will cause the material to split or tear from the edge. A poor blanking operation will greatly reduce this permissible tensile elongation and the depth of flange which can be generated.

2.3.4 HOLE EXPANSION

In the strictest terms, hole expansion (Figure 2-8) is a type of stretch flanging in which the bend line is a complete circle. Accordingly, another name for the operation is hole flanging. The reason this operation is given a separate classification is twofold. First, hole expansion is a very distinct operation in many stamping plants and is therefore uniquely termed on blueprints. Second, as will be explained later, hole expansion is a popular test used to evaluate sheet metal forming capacity. The same comments made above concerning stretch flanging are applicable here. The strain state shows the tensile major axis of the ellipse in the circumferential direction around the hole. There is a compressive minor strain near the free edge of the hole and a zero minor strain close to the bend line.

2.3.6 EMBOSSMENT
The simplest of the forming operations is embossment. The operation is performed on a localized area such that the remainder of the blank is large compared to the deformation zone. This means that the blank is considered to be of infinite size and no metal flows from the blank into the deforming zone. Embossments are generally divided into three types (Figure 2-9). These are: 10 beads and ribs, 20 offsets, and 30 decorative (e-2). The beads and ribs are characterized by a long, narrow depression (Figure 2-9a) which may be straight, circular, or combination. The offset is the displacement of a large area of metal (Figure 2-9b), in which the edge (B) is one-half the width (A) of the bead. The third category is any combination of beads and offsets used for decoration; a most common example is embossments of letters and numbers on automotive license plates. A common characteristic of all embossments is the displacement of metal without a reduction in sheet thickness in the flat section of the offset (E-2). Embossed designs are visible from both sides of the sheet, though the relief is reversed. Thus, an embossment is different from a stamp in that a stamp depresses sheet thickness from only one surface. Grids placed on embossments would show a major tensile strain across the radius and no strain (plane strain) along the radius. All other metal areas remain unstrained.

2.3.7 BIAXIAL STRETCH


Biaxial stretch forming is the only operation in which both the major and minor strains are tensile. A common example is the penetration of a hemispherical punch into an infinite sheet or a sheet effectively locked by beads (Figure 2-10). Because no metal can flow into the die area from the flange (binder) area, all deformation is restricted to the metal initially within the die opening. This forces the metal to stretch in all directions. The ellipse, therefore, is elongated in both directions for a (+ by +) strain state.

2.3.8 DRAW
In the automotive press shop, most dies are called draw dies because the metal is drawn into the die cavity. In reality, most of the deformation is biaxial stretch over the punch or bend-and-straighten metal flow from the flange. Drawing, sometimes known as cup drawing, radial drawing, or deep drawing, has a very specific set of conditions which differentiate it form these other operations. The most common drawing operation is the formation of cylindrical cup (Figure 2-11). The blank is drawn into the die cavity by the action of a flat bottom punch. Deformation is restricted to the flange areas of the blank. No deformation occurs under the bottom of the punch the area of the blank originally within the die opening. This is the exact opposite of stretch forming. The unique character of deep drawing is the deformation state of the flange. As the blank is pulled toward the die line its circumference must be reduced. This reduction in the circumference generates a

compressive stress in the circumferential direction, resulting in a radial elongation as the metal is extruded in the opposite direction. Here is an appropriate time to differentiate between different stress states creating identical strain states. This very important differentiation is often overlooked, yet changing stress states from tensile to compressive can mean the difference between failure and success. Figure 2-12 illustrates the difference. A circle can become elongated by two methods. A tensile stress (force) can be placed on axis A-A which will result in a contraction along axis B-B. Alternatively, axis B-B can be compressed and the metal will elongate along axis A-A. However, the tensile stress of the first example is much more damaging than the compressive stress of the second case. A common analogy would be generating a ribbon of toothpaste from a toothpaste tube. Pulling the ribbon of toothpaste out of the tube by a tensile pull would not be successful because of tensile necking. However, compressive extruding of the toothpaste is very simple. Another example of different stress systems for the same strain state is the through-thickness compression of a sheet (Figure 2-12b) compared to a balanced biaxial tension in the plane of the sheet.

2.4 Subsequent Forming Operations


Many parts are formed in a sequence of forming operations as opposed to a single operation. The primary reason for multiple operations is that the severity of the forming is too great to be accomplished in a single operation. A prime example is the restrike operation (Figure 2-13). Attempting to stretch metal into a die cavity by stretching over a very sharp radius punch (a) would result in failure. Instead the metal surface area or length of line is generated by stretching over a generous radius punch. A restrike (b) then redistributes the metal into the desired configuration without an additional tensile increase in the length of line. Another common subsequent forming operation is redrawing. Limits are imposed on the blank diameter which can be drawn into a cup of a given diameter. Should a deeper cup be required, an intermediate diameter cup is drawn first. This cup is then redrawn in one or more subsequent stages to achieve the final diameter and height. In some case the cup is ironed to obtain additional cup height. These subsequent forming operations are well described in textbooks (E-2, S-3). It is important to note that the limiting stage, and the most critical forming operation (E-1, K-24). Thus, the importance attributed to the first forming operation is well founded.

2.5 Interaction of Forming Modes


Most stampings are composed of one or more primary forming modes. The two shapes shown in Figure 2-1 are reproduced in Figure 2-2 but the geometrical designations have been replaced by the respective forming modes used to generate the geometrical shapes. Once the forming modes have been identified, the analysis of the stamping can begin. One such method is the length of line analysis technique. Here the required length of line is analyzed for each of the directions and zones in the stamping. An example is shown in Figure 2-14. In this complex stamping, note that the wall geometry of line 1 is generated by a radial or cup drawing mode. This is determined by examining the plan view of the stamping (Figure 2-15). Note that the four corners can be combined to form a complete cup, which is created by radial or cup drawing. By contrast, the wall geometry of line 2 is generated by a bend-and-straighten operation. Thus, it becomes important to identify known forming modes within a stamping. Each of the forming modes requires different material properties for optimum formability. Both design parameters and material properties will vary greatly often in the opposite direction depending on the forming mode (Discussed in Section 3). After the first stamping is formed, a wide option of secondary forming operations is available to change the shape created by each of the initial forming modes. For example, trimming and reverse flanging of the

initial flange may present problems. The bend-and-straighten section remained at initial blank thickness. However, the metal in the corner cup draw has increased in thickness and may be up to 40 percent thicker. Compensation for this change in thickness must be made in the clearances of the trim die and the reverse flange die. Even more complex, these forming modes are interactive and are constantly changing in response to a large number of variables. For example, an increase in blank width to provide an additional flange after trimming for welded requirements will restrict metal flow from a binder (holddown) area and thereby increase the depth of the stamping required by stretching over the punch. This may drive the deformation over the punch into a failure condition. Increasing the die radius can reverse this trend by allowing easier flow of the extended blank into the die cavity. However, if this die radius now is too large for the final part, a restrike operation will be required to sharpen the radius to the required (not desired) dimension. These are the interactions which must be considered by the simultaneous engineering team. Unfortunately, an increase or decrease in interface lubricity with respect to the surface characteristics of the incoming steel can inadvertently cause the same reaction in the die and cause part to part variability. Thus, the first task at hand is to understand the forming modes and how they interact to provide the desired part (and therefore stamping) shape.

2.6 Auxiliary Operations


These operations are performed on the incoming metal, during intermediate stages, or on the final part to make it conform to blueprint specifications. These operations include trimming, flanging, piercing, etc. and are well described in various textbooks (E-2, S-3). An important consideration with these operations is that they are often performed on metal which is in the work hardened state and therefore has reduced formability.

2.7 Illustrative Example


A common part is the rectangular, flat bottom box (Figure 2-16). At first glance, view a), this might be called a pure deep drawing operation according to traditional, current press shop terminology. The crosssectional view b) is identical to that found in deep drawing. However, analysis of the plan view c) indicates a combination of deep drawing and bend-and-straighten operations. If the four corners are removed and reassembled by themselves, a deep drawn, flat-bottom, cylindrical cup is created. Thus each of the four corners is one-quarter of a pure deep drawing operation. Examination of the remaining four straight-sided sections shows these areas to be formed by bend-and-straighten operations. Elements move inward from the flange without any deformation or strain until the radius is reached. The (+ by -) strain characteristic of deep drawing is absent. In passing over the die radius into the pan wall a (+ by 0) deformation takes place. In reality, the compression action of the corners forces excess corner material into the straight flange areas (Figure 2-16d). This reduces the compression strain in the corner and permits a deeper box to be generated than that predicted by pure deep drawing formulas. Note the different characteristic patterns of buckles or wrinkles in the corner and side segments. In draw zones the wrinkle lines converge at the die radius. In bend-and-straighten areas they often will diverge at the die radius. Note the wrong labeling of draw beads used to restrain metal flow. Draw beads are not placed in draw regions of the stamping; in draw regions metal flow from the flange is to be encouraged, not restricted. Metal flow restraint often is needed in the bend-and-straighten areas to control excessively rapid metal flow and to impart a small increment of stretch to eliminate loose metal. Therefore, draw beads actually should be renamed to be stretch beads (E-1).

The part sketched in Figure 2-17 previously would have been classed as a drawn part using the old press shop terminology. Yet analysis of this part based on the circular grid analysis system would indicate that no areas of deep drawing are present. Zone A is generated by a stretch operation. Zone B is generated by a simple bend-and-straighten operation. No compressive minor strains, required for deep drawing, are present in this part. A good composite stamping which illustrates the various strain states is a round-bottom, cylindrical cup (Figure 2-18). The top portion of the cup is formed by stretch forming over a hemispherical punch. The strain states are (+ by -) or tensile-compressive. Between the two regions the minor strain becomes zero. This zone of (+ by 0) or plane strain is considered to be the dividing line between stretching and deep drawing. This zone also is the restrike or die impact line observed on most stampings. Unfortunately, we are currently in a transition period where both the old press shop jargon and the new terminology systems are used simultaneously. This creates a tremendous problem in communication. Persons discussing a problem and using the same terms can mean directly opposite forming operations. The greatest offender is the term draw die. It is, therefore, important that this report begin with the new definitions of the forming operations, as this entire report will be based on this revised terminology.

2.8 Summary
Sheet metal forming is a complex interaction of numerous forming modes. Prediction of sheet metal forming limits or analysis of forming failures depends on accurate assessment of these forming modes. A number of problems are encountered with this concept. DEFINING THE FORMING MODES. Some method of defining the modes of deformation is required. Geometrical definitions are not capable of sufficient discrimination. For example, a wall of a stamping could be generated by several different forming modes-some of which are more severe of damaging than others. Therefore, the actual deformation of the metal, combined with the stress state necessary to create the deformation, currently is the best framework for describing forming operations.

NAMING THE FORMING MODES. Once the forming modes are described in terms of the deformation patterns and deformation histories, proper naming of the forming modes naturally follows. The problem now becomes one of conflict with traditional press shop jargon, historically generated from a different reference base. The most common conflict is to call the first die in a press line a draw die when very little or not drawing actually is created in the die based on the engineering definitions discussed in this section. IDENTIFYING THE MODES WITHIN A GIVEN STAMPING. Few practical stampings are generated by a single forming mode. Instead, numerous forming modes are needed to create the common, complex stampings. These modes interact and compete with each other during the forming cycle of the stamping. Thus, some method of identifying the various segments of the stamping with the correct forming mode is required. DESCRIBING FAILURE IN TERMS OF A SPECIFIC FORMING MODE. The corrective action for sheet metal forming failures depends on the specific mode within which the failure occurs. To say the stamping failed is no longer sufficient for effective correction of the problem. The forming mode associated with the failure must be identified. This will be discussed in more depth in subsequent sections. ANALYZING CHANGES IN FORMING MODES. The forming modes, as well as the location of failures, change during the production life cycle of a stamping. These changes are gradual cover time with discrete steps often occurring when the tooling is replaced in the press. These changes need to be detected and tracked for effective forming analyses.

An important first step to improved sheet metal forming analysis, therefore, is correct definition and identification of the various forming components in all stampings. Figure 2-1: Schematic of a stamping with a combination of geometries.

Figure 2-2: Schematic of a stamping showing different possible forming modes.

Figure 2-3: Schematic (a) showing the action of a Vdie, Udie, and a wiping die. The strain state in bending is shown in (b) (E-2).

Figure 2-4: Bend-and-straighten deformation occurs as the metal moves from the flange to the wall.

Figure 2-5: A large component of stretch added to a small bending strain produces two different

tensile strains on the convex and concave sides.

Figure 2-6: The tip of the stamping in a die can cause a change in the type of deformation.

Figure 2-7: Shrink flanging in the top sketch (compression) is compared to stretch flanging in the

botom sketch (tension).

Figure 2-8: In hole expansion, the material at the edge of the hole elongates in the circumferential direction and contracts in the radial direction, thereby approximating the strain state in a tensile test.

Figure 2-9: Two configurations are possible for an embossment.

Figure 2-10: Stretch forming in which no deformation is allowed in the blank area and all

deformation occurs in the die opening over the punch.

Figure 2-11: Deep drawing in which no deformation occurs over the bottom of the punch and all deformation is restricted to a tension-compression deformation in the flange of the blank.

Figure 2-12: Identical strain states can be obtained from either tensile or compressive stress.

Figure 2-13: Second forming operations are more successful when the length of the line is generated in the first operation over generous radii.

Figure 2-14: Line analysis of a complex stamping.

Figure 2-15: Plan view showing cup draw segments in the corners.

Figure 2-16: Most "deep draw" forming operations are primarily composed of bend-and-straighten segments. Only the four corners reflect a true drawing deformation.

Figure 2-17: While this stamping looks like a deep drawn part, circle grid evaluation reveals

stretch forming (A) and ben-and-straighten (B) operations.

Figure 2-18: A round-nosed cup is composed of balanced biaxial stretch at the pole, biaxial stretch over the remainder of the punch nose, a line of plane stretch (- by 0) seperating stretch and draw, and cup drawing.

3.1 Introduction
Historically, the progression of an automotive sheet metal stamping from conception to production has been a segmented series of events. Styling, part design, material selection, die design, die build, die tryout, and part production have been performed in a sequential manner. Interactions between adjacent stages have been minimal at best. Manufacturing feasibility discussions between part designers and production staffs have been almost nonexistent. For example, die designers have had little input into the design of the part and rarely did they interact with the press room tryout staff. The activities of each segment in the sequence have been conducted within its own sphere of work by its own group members. Interaction among the different groups has been limited. This problem is improved by the recent growth of simultaneous engineering (A-34). Here all groups involved in the design to production chain meet early in the design stage to provide input into the design of the part. For example, the production staff may suggest a minor part change which will radically reduce the amount of offal. Ideally, the simultaneous engineering concept even brings specialists from the

material, lubricant, and other outside suppliers into the initial design phase, where major design changes can be made most easily in a cost/time effective manner. Implementation of the simultaneous engineering team is an excellent step forward. However, the input of each team member generally is limited to his personal knowledge most likely gained through years of trial and error experiments. This is especially true in the area of sheet metal formability. Lacking here are master design guidebooks, which detail the specific limits for different types of sheet steel for each of the forming modes described in the proceeding Section2. Few sheet metal formability design limits are available. This is especially true for the automotive industry which forms an infinite variety of complex designs from a large variety of coated and uncoated steels. Some of the other industries have less of a problem. For example, the container industry produces many billions of cans which are geometrically identical (drawn and ironed or draw-redraw), symmetrical in shape, restricted to limited steel grades and coatings, and related to a historical design which has been fine-tuned for decades. Design rules or limits for automotive stampings which do exist are fragmented among many sources. A few textbooks are available, such as the important text by Eary and Reed (E-2) and others (A-4, A-10, A12, A-13, A-14, A-15, D-14, J-10, K-28, L-3, S-11, S-14, S-29, S-30, W-7, C-1). The AISI has attempted to tabulate some of these design rules in recent publications. The first was Sheet Steel Formability (A-12). A section on Formability is included in the Automotive Steel Design Manual (A-10). Some books use the case history method to present design/forming limitations (A-14). The references which are available are extremely limited and do not provide any coherent system for data acquisition. Ideally, mathematical modeling of sheet metal formability (Section 8.2) is intended to replace the need for manuals or personal knowledge of sheet steel design rules. However, this capability will not be available to every type of designer in all automotive companies (small component suppliers as well as the major companies) for many years. The design manuals are needed in the interim. Finally, a the mathematical modeling systems are developed, the design manuals are an effective method to test the predictive capabilities of the new mathematical modeling systems. Several foundations are required to support the concept of a sheet steel design manual, however. These include a common language (Section 10), a general definition of the forming system (Section 1), a uniform description of the component forming modes (Section 2), and an understanding of the contributions made by components of the forming system other than design (Sections 4-6). This section will highlight some of the areas where design rules are available and areas which lack significant guidelines.

3.2 Blank Development


The stamping operation begins with the creation of a blank. The blank is created by either shearing or blanking. In shearing a straight cut is made across the coil width to form a square or rectangular blank. In blanking a contoured blank is created. The perimeter of this contoured blank is composed of combinations of straight and curved segments. While the shearing operation is the easiest to perform, it may be very wasteful of metal (Figure 3-1a). Nesting of the blanks can reduce the metal which must be removed by trimming (Figure 3-1b). This trim material is called offal or engineered scrap. It is metal intentionally wasted from the initial blank on every stamping. Sometimes large segments of offal can be reapplied on smaller parts but careful study of the economics must be made. The costs of steel collection, storage, reapplication, accounting, scheduling, etc. may outweigh the savings generated by the amount of the scrap metal actually used. A primary goal of contouring, however, is to provide a nested pattern of blanks which will minimize offal.

Another reason for the contouring of blanks by a blanking operation is to match the blank perimeter to the perimeter of the die opening. This encourages more uniform metal flow into the die cavity, prevents excess metal buckling in the flange, and reduces the drag of extra flange metal behind critical zones. A third reason for contouring the blanks is to create the final flange/part contour in the blank in order to eliminate a trim after the forming operation. Obviously, the contour of the blank with respect to forming may be in conflict with the blank shape for ideal nesting. The shearing/blanking operation can be performed in several ways. It may be done in material receiving, where coils of steel are blanked and the stacks of blanks shipped to the press line. The blanks may be made at the head of the press line as the coil is unwound into a washer, blanker, oiler combination. Finally, the blanks may be created in the die at the start of the forming stroke. All the cutting operations detailed above involve the same basic theory of cutting sheet metal (E-2, J-10). The cutting occurs by a combination of metal penetration and actual fracture of the metal. A critical factor in both the visual appearance and the residual ductility of the cut edge is the clearance of the cutting knives. For low strength, low carbon steel an aim of ten percent clearance is used for automotive body panel stock. The most important stamping consideration for cutting is the residual ductility remaining in the cut edge (M-39). Figure 3-2 shows the residual stretchability as a function of burr height. Here the burr height is the measure of damage inflicted during the cutting operation. More specific information is contained in Figure 3-3. Here the hole expansion values are provided for different strength steels as a function of the quality of the blanked hole. The hole expansion test simulates closely the permissible elongation of a blanked edge. Similar data have been presented in terms of permissible elongation for given grades and thicknesses of steel. This is a fruitful area for additional work. While standard rules of thumb are useful for example shear clearances of 10 to 14 percent of metal thickness they can not always be met. Some secondary blanking/trimming operations which occur between forming stages are constrained by the geometry of the stamping relative to the tooling. In these cases, additional damage to the blanked edge is unavoidable. Here it becomes important to know the permissible strain level and strain distribution capabilities of the resulting blanked edge so that subsequent forming stages do not result in failure (Figure 3-4). Such information may have to be inserted into mathematical modeling systems in tabular form, as calculation of permissible residual formability after blanking may not be possible from fundamental plasticity theory. Extensive work by Hugo H-4, H-41) on hole punching provides some insight into the subject of blanking, since hole punching can be viewed as circular blanking. His data are summarized in Figure 3.5. The punch loading curve (Figure 3-6) shows how punching (blanking) energy, measured from area under the load-penetration curve, can be reduced by using larger punch-to-die clearances. Generally, process conditions for the minimum energy are optimum. Additional energy in consumed in redundant work, excessive deformation, heat loss, etc. Much of the additional energy consumed is translated into damage of the sheet metal. If the stretching required to make part print exceeds the residual stretchability after cutting, several avenues are available for restoring the edge stretchability. One is improving the quality of the original cutting operation. A second is an additional cut of higher quality. A third would be a shaving or milling of the cut edge before subsequent tensile straining. A fourth is heating of the cut edge to reduce the effects of the cutting operation; this has been described as thermal deburring. If these steps are not feasible solutions, the stamping/part will have to be redesigned to reduce the required stretching on the blanked edge.

3.3 Generating the Stamping Wall


The walls of the stamping may be generated by one or more of the forming modes described in Section 2.3. The walls are composed of the sides and the corners (Figure 3.7). The side walls are further subdivided by forming modes flanging and bend-and-straighten. In like manner, the corners are subdivided into flanging and drawing.

3.3.1 SIDE WALLS


3.3.1.1 Flanging
The walls of some stampings are created by flanging. While flanging may be done on a press brake, the same deformation mode occurs with a wiping die or a die without a blank holddown ring or binder. The flanging of the sidewall can be further subdivided into straight line flanging and curved line flanging. STRAIGHT LINE FLANGING - This flanging operation simply is a bending operation where the sheet steel is folded (bent) over a radius. Deformation occurs only in the plastic hinge over the die radius with no strain along the axis of the bend. Bending is different from other forming modes in that a severe strain gradient is developed from one surface to the other by the act of bending. The maximum tensile strain is found on the convex surface of the stamping, while the maximum compressive strain is found on the concave surface of the stamping. Somewhere within the sheet metal is a neutral axis which does not change length. A common design assumption is that the surface area of the bend at the neutral axis, and therefore the thickness of the bend, does not change. The convex tensile and concave compressive stresses tend to magnify the elastic recovery or springback reaction when the forming loads are removed. Harder materials, thinner sheets, and larger bend radii increase the tendency for springback. Methods for controlling springback include overbending and stretch bending. Overbending does not reduce springback but simply adds an increment to the original bend such that the original bend angle minus the springback will equal the design angle. Stretch bending actually reduces the amount of springback. By adding a through thickness tensile component, the tensile/compressive stress gradient is reduced. Springback is discussed in detail in Section 8.3. The radius of a bend in sheet metal forming generally is specified in terms of thickness of metal being bent. A sheet with a thickness of 0.030 inches (0.75mm) bent over a radius of 0.12 inches (3mm) would have a 4t bend. Bends therefore can vary from zero thickness to infinity (no bend) at one location and the bending can change from one location to another, even reversing the amount and direction of the bend. For automotive stampings the metal thickness of cold rolled steel is generally in the range of 0.020 inch (0.5mm) and 0.070 inch (1.8mm). The outer body panels will cover a smaller range from 0.025 inch (0.6mm) to 0.04 inch (1.05mm). The bend radii for most of these stamping should be between 3t and 9t. Larger radii up to 25t are used if the design requires a rounded appearance. Too sharp a bend will cause excessive tearing, especially if there are tensile forces associated with the bend. These sharper bends usually require a restrike operation. Too large bend radius introduces the possibility of excessive springback and the necessary larger spacing needed between the bend dies can cause loss of control of the metal being bent. Bends are most frequently made to a 90 degree angle by the vertical movement of a punch into a die opening. A hinge die can be used to overbend beyond 90 degrees. Further bending and flattening to form a hem can then be accomplished by a subsequent punch action against a backing plate.

A system for predicting bending limits, which will serve a function parallel to that of the forming limit diagrams for membrane-type deformation, must be based on the maximum allowable strains generated in the outer (convex) surface of the bend. Studies have shown (Figure 308) that the strains on the outer surface depend upon the bend angle and the ratio of bend radius to sheet thickness (r/t). Interestingly, outer fiber strains do not change drastically for different r/t ratios until the bend angle exceeds 50 degrees. This is shown in Figure 3-9, which is a replot of the data contained in Figure 3-8. Since the outer surface strains are a direct function of the radius curvature of the bend (Figure 3-10), the different r/t bends must generate different radii of curvature in a free bend (press brake) operation even for constant punch radii. Even more significant in Figure 3-8 is the constant peak strain observed for increasing the bend angle beyond 90 degrees for r/t equal to 1.67. This is consistent with an earlier graph (Figure 3-11) by Keeler (K-24) which shows that a peak bend strain is reached after which further bend depth (increasing bend angle) does not increase the peak strain but causes the lower peak strain to be reached at an earlier bend angle. This agrees with the two strain distributions shown in Figure 3-12 and the series of strain distributions in Figure 3-13. The above studies indicate that the outer fiber stain is a measure of bend severity and depends upon the geometry of the bend. The studies also show that the most critical stage in a bend operation is not at the final 180 degree angle but somewhere earlier in the bending operation. For satisfactory bends and larger, Figure 3-8 suggest the critical angle is approximately 120 degrees. The other half of the problem is to define the maximum allowable strain the outer surface can withstand. One study (K-24) suggested the maximum outer fiber strain a metal can withstand is related to the total elongation of the metal. This sounds feasible but no definitive data can be found. Figure 3-14 shows poor correlation between bendability and percent total elongation and better correlation between bendability and percent reduction of areas. However, traditional bendability data are not sharply defined by outer fiber strain discussed previously but by a pass-fail system at 1t and flat bend. The same pass-fail system is used to show the effect of stringer inclusions in Figure 3-15 for different yield strengths. Relationships developed form this pass-fail system should be used with extreme caution, as should various bend allowance tables based on the 1t, 2t, etc. system. A bendability definition which is better than the pass-fail system is the length of the longest edge crack (K55). This test combines outer fiber strain and the strain limits of the material. Using this criterion, an important graph (Figure 3-16) shows the effect of sulfide shape control during bending. The graph shows reduced bendability of an 80 ksi (550 MPa) steel compared to a 50 ksi (345 MPa) steel when neither steel has inclusion shape control. With inclusion shape control, both steels have equal bendability at a level significantly better than the 50 ksi (345 MPa) steel without inclusion shape control. Bending pressures (free bend) for various metal thicknesses and die openings are recorded in the literature for steel (S-13). In general, however, the loads used for bending operations (including roll forming) will be proportional to the yield strength times the square of the metal thickness (B-18). A gage reduction accompanying a strength increase can result in lower press loads. Conversely, thickness increases when changing to aluminum form 1008 steel at approximately the same yield strength can greatly increase press loads required for bending because of the thickness squared relationship. In a wiping die, bend forces are minimal. However, pad holding pressures should be approximately 10 times the required punch pressure to avoid sheet slippage and recoil (E-2). Flanging die materials need more attention for high strength steels than aluminum as higher loads will be encountered. Wear and galling resistance are of increased importance for both high volumes and difficult forming conditions.

While die materials such as T-15 offer perhaps the best performance, they are more difficult to heat treat and weld than the oil or water hardening steels (B-18). Experiments for a wide variety of metals (Y-3) have shown relationships between minimum r/t ratios and total elongation (etot) and reduction in area (%RA). These results are shown in Figure 3-17 and can be translated into design formulas, such as:
Rmin = Rmin = 50 tot 50 % RA
e

- t/2 -1t

where Rmin is the minimum bend radius for successful bends. The analysis becomes complicated when laminated (steel-core-steel) sheets are evaluated (S-19, Y-14, Y-6, M-17). Flanges are short vertical bends at the edges of a panel or surrounding a hole. When bending a flanged inside corner, the corner radius should be at least 4t (A-10). If an outside corner is to be formed with a flange, the minimum radius of the corner should be 5t and the angle of the corner no sharper than 60 degrees. In hemmed corners where the metal of the flange is folded back against the sheet, the minimum allowed corner radii are increased to 24t for an inside and 7t for an outside corner, again no sharper than 60 degrees (A-10). Hem flanges are used to strengthen the edges of sheet metal parts, give a smooth rounded edge to a part, or to provide hidden joints. They can be either concave or convex, but in either case problems of too little metal for an inside flange, or too much metal for an outside flange, must be handled during the bending process. This is accomplished by limiting the width of the flange, cutting notches in the corner flange metal to reduce the amount of metal to be bent and stretched or shrunk, or designing offsets to take up excess metal. Offsets are displacements of a few metal thicknesses, similar to those used to form license plate numbers. For corner flanges, the offsets can be considered designed-in wrinkles. Edges that must be strengthened further than is possible by hemming are curled. CURVED LINE FLANGING Curved line flanging has an additional component of strain added to the deformation. As the blank is folded, the blank must elongate or shorten along the length of the wall to conform to the geometry of the wall. This deformation increases as the vertical distance from the bend axis is increased; the further away the element is located, the more the length of line must change. For flanging on a convex bend axis, the length of line must become shorter. This forming mode is called shrink flanging. Because the stresses are compressive in shrink flanging, breakage is not the failure mode. However, shrink flanges tend to generate buckles and loose metal. Careful control of punch/die clearance is required to produce a clean flange (N-8). In contrast, a concave bend axis requires the length of line to increase. This forming mode is called stretch flanging. A stretch flange will generate edge cracking and tearing if the stretching limits are exceeded. The simplest analysis of stretch flanging is to calculate the increase in the length of line assuming that the stretch flange is a segment of a circle. These elongations for steel are then compared to the permissible hole elongation limits (M-12, F-4 and J-12). These limits are conservative, however. Unlike true hole expansions, where the entire perimeter of the hole is subjected to the same elongation, the adjacent metal in a stretch flange often is undeformed or may even be in the shrink flange mode. Adjacent areas to the stretch flange may be able to feed metal into the stretch zones and reduce the required elongations.

The hole expansion test utilized in the laboratory is described in detail in Section 4.3.4.1. This test closely duplicates production stretch flanging. Typical hole expansion data are given in Figure 3-3. These tests show that stretch flanging limits increase with increasing sheet thickness, work hardening exponent (n), total elongation, normal plastic anisotropy (rm), and minimum plastic strain ratio (r) and decrease for increasing yield strength, tensile strength, and hardness. The quality of the blank edge plays a major role in the stretch flanging limits (Figure 3-3). Any damage to the edge of the blank, either from the creation of the edge or subsequent edge damage, will drastically reduce the edge elongation and therefore reduce the design limits of the forming operation.

3.3.1.2 Bend-And-Straighten
The previous discussion assumed that the walls of the stamping were created by metal swinging in free space while deformation was restricted to a plastic hinge. In reality, most stamping walls are created by metal being pulled from the binder or holddown area where the sheet metal is restricted to sliding between the upper and lower die segments (Figure 3-18). The bend-and-straighten operation is a rather simple forming mode. Metal flows in a straight line path through the binder zone. The metal then bends to conform to the die radius, flows over the die radius, and then straightens to conform to the die wall. Even with ideal metal flow from the binder, the wall of the stamping will have undergone cold work during the bending and unbending sequence. This cold work reduces the capacity of the steel for subsequent forming operations. The amount of cold work, and therefore the severity of this operation, depends on the ratio of the bend radius to the sheet metal thickness. For die radii less than 4t a severe tensile strain is generated on the convex surface during bending and on the concave side during straightening. Because the tension on the concave side follows compression (which work hardens the metal and depletes useable formability), the concave side of the bend-and-straighten side is the more severe forming operation. On the other extreme, the die radius should not be more than 10t. Greater radii than that creates a band of unsupported metal between the die radius and the punch radius. The ideal die radius would be 6t and 8t. Any tension created in the binder area can add a stretch component to the pure bend-and-straighten operation. This restraint can be created by additional binder or holddown pressure or by the insertion of draw beads into the binder surface. Beads are placed in the bend-and-straighten areas to restrict metal flow so that the blank will not run into the die cavity too rapidly. This restriction will create an added stretch component. Therefore, strictly speaking, the draw beads would be more accurately identified as stretch beads.

3.3.2 CORNER WALLS


The corners of the wall structures present special problems in forming. Sheet metal stampings tend to be designed with rather sharp radii in the corners.

3.3.2.1 Flanging
Most corners formed by flanging are equivalent to the shrink flange described above. The length of line must be decreased, thereby creating opportunities for buckling. The higher the corner wall, the more the line length must decrease. Depending on the forming mode of the sidewall segment adjoining the corner, some of the excess metal will be forced into the sidewall areas, thereby reducing the compressive stress. Minimum punch to die clearance is used to iron out the buckles as the corner metal enters the die cavity. Excessive buckles can cause the metal to fold over on itself; this can jam in the punch-die opening and cause the bottom of the stamping to rip open.

3.3.2.2 Drawing

The drawing mode for generating wall corners is similar to the bend-and-straighten mode for generating sidewalls (Figure 3-19). In both cases the sheet metal is restrained by the upper and lower segments of the die. The metal slides through the restraint towards the die; this motion is designated as radial motion. Unlike the bend-and-straighten operation, however, the sheet metal is forced to undergo a circumferential compression. It is this circumferential compressive strain that is unique to drawing and places special property requirements on the sheet steel (see Section 4.1.4.8 Anisotropy). The stamping in Figure 3-19 shows four corners separated by four sidewalls. If the four corners are removed and joined together, a cylindrical cup is formed. In practice, some production stampings are all wall corners without sidewalls. An oil filter and power steering pump housing are examples. The cylindrical cup can now be related to the laboratory cup drawing test in an attempt to develop some design limits. Most of the design data in the literature is available for the radial or cup drawing mode of deformation. Historically, this operation has been uniquely identified, easily reproduced in the laboratory, and amenable to systematic investigation. For these reasons, successful correlation between laboratory experiments and actual press shop behavior has been obtained. Of all the forming operations, cup drawing in the most predictable and has the most known design parameters. The axi-symmetrical nature of cup drawing lends itself to mathematical analysis. As a general statement, however, press shops consistently violate cup drawing formability limits which have been published for decades (E-1, S-3, J-8, H-9). Two measures of cup drawing severity are used, both of which are mathematically interrelated. The most common is the Limiting Drawing Ration (LDR = D / d) which is the ratio of the maximum blank diameter (D) which can successfully be drawn into a cup of punch diameter (d). The other measure is the percent blank diameter reduction given by:
% Reduction = D-d D 100 100 - % Reduction x 100

LDR =

x 100

A common reference point is a 50 percent reduction equals an LDR of 2, which is a common value for rimmed steel. A similar formula is available (E-2) for an approximate calculation of cup height.
h= 4d For an LDR of 2, h= 3d 4 D d

The most important unknown in cup drawing, therefore, is the maximum severity of deformation (largest blank diameter) that can be formed into a cup of given punch diameter; this is measured by Limiting Draw Ratio or LDR (K-24, W-19, W-20, C-15). LDR can be readily measured in the laboratory. This test is described in detail in Section 4.3.4.10. Here the LDR results are shown to depend on: plastic anisotropy ratio (rm) of the steel, the sheet thickness, punch and die geometry, test speed, steel temperature, lubricant, holddown loads, surface topography, and many other parameters. For this reason, the effect of various steel and forming process parameters can be assessed in the laboratory and the results ranked. However, the interaction of the various parameters prevents the determination of an absolute value of the LDR or permissible cup height which can be translated to a simple production cup draw, much less a complex stamping where drawing is restricted to only one corner of the stamping.

An example of this inability to transfer absolute numbers to production stampings is illustrated by Figure 3-20. Here the desired punch radius is shown as a function of sheet thickness. If the punch radius becomes too sharp, the localization of strain over the punch radius due to stretching of the material becomes too great and the strain level the material exceeds that permitted by the Forming Limit Diagram. Sheet thickness affects the punch strain in two ways. First, the level of the Forming Limit Diagram (permissible ceiling strain) decreases with decreasing with decreasing sheet thickness. Second, the reduced sheet thickness reduces the bending strain; this in turn reduces the bending load component of the total load which can be transmitted to the deforming flange. To maintain a total forming load in the flange, the stretch load and therefore the level of the stretching strain must increase. However, the data in Figure 3-20 were generated under one set of test conditions. Increase the coefficient of friction of the punch radius interface and additional stretching will cause the curve to shift upward. Reduce the coefficient of friction and the curve will shift downward. In the extreme, welding of the steel to the punch radius will permit infinitely sharp radii. Thus, while trends can be determined, absolute design parameters can not be determined without knowing the interaction of other parameters. If a deeper cup is required than allowed by the LDR, one or more redraws must be employed. The interrelationship of blank diameter to cup height and flange width is shown in Figure 3-21. Note the decreasing D / d reductions allowed for each additional redraw. Eary (E-2) suggests the following schedule of redraws if t / D>0.5: 1st draw = 50% reduction (as large as possible) 1st redraw = 30% 2nd redraw = 23% 3rd redraw = 19% 4th and subsequent redraws 10% This rapidly decreasing redraw allowance does not make large numbers of redraw stages profitable. Therefore, each early stage should be made as severe as possible without breakage. For geometric reasons and maximum reductions (Figure 3-21), no flange should remain on the cup until the last possible forming stages (H-34, M-1). The anisotropy which leads to high rm usually creates a high r for common 1008 steel (H-42, W-22, G40). This in turn creates increased earing, which reduces the usable depth of the cup after trimming. The relationship of ear height and r is shown in Figure 3-22. The percent earing also increases as a more severe draw ratio is used. This is shown in Figure 3-23 for aluminum, but similar curves are available for steel. The percent earing can be reduced by an ironing operation (Figure 3-24), which modifies the normal thickness gradients around the circumference of the can wall. Eary (E-2) has published a well-documented list of variables and their influence on punch loads: a. Punch no effect on punch load b. Die radius 35 percent decrease for increase from 2t to 10t c. Lubrication 26 percent decrease for change from no oil to heavy oil d. Blankholder force no effect for a design of 1/3 of punch force e. Percent reduction 100 percent increase for an increase from 37.5 to 50 percent f. Depth of cup increase from zero to maximum for initial 1/3 stroke g. Speed no effect until punch speeds exceed 120 feet/min. (36.6 m/min.) Force requirements also were predicted by Korhoen and Sulonen (K-51). The influence of the work hardening exponent is shown in Figure 3-25. The maximum drawing stresses increase as draw ratio and the work hardening exponent increase. The limiting draw ratio is reached when the maximum drawing stress is equal to the tensile strength of the cup wall. These limits are shown

by the downward arrows and are approximately constant at a drawing ratio of 2.2. This agrees with various other theoretical studies (K-14, M-19). Press loads for higher strength steels increase as yield strength increases. This increased load is balanced somewhat by the usual accompanying reduction in sheet thickness. Die design for the higher strength steels, however, should follow the practice for thick, low carbon steel drawing (B-18) and specify cast steel punches and D-2 steel binder inserts. An important problem in deep drawing operations is the adjustment of the blank holder force to prevent blank wrinkling (N-3). At higher punch speeds, the flange forms a number of wrinkles equal to the number of ears which would finally form on a drawn cup. Blankholder pressure required to suppress wrinkling increases as the r value decreases. Blankholder pressures increase with increasing r, suggesting the minimum r value is a critical parameter. Havranek (H-7) found that a higher strain hardening exponent (n) leads to increased wrinkling, which has to be compensated for by an increased blankholder load and decreased die profile radius. He noted that wrinkles occurred readily in conical cups produced from HSLA steels due to its higher flow stress and lower r values (H-8). A cup drawing limit diagram is now commonly found in the literature (Figure 3-26). The permissible range of blankholder force is presented as a function of both wall wrinkling and punch nose fracture limits. Deeper cups require a narrower range of permissible blankholder forces. These cup limiting diagrams are influenced by tool geometry (Figure 3-27) and steel properties (Figure 3-28). Another form of these splitting/wrinkling limits has been detailed by Stine (S-39) Again, Eary (E-2) has defined potential wrinkling problems as a function of t/D (sheet thickness to blank diameter ratio).
When t/D percentage is 0.50 or less Wrinkling is very severe and compressive loads must be reduced. A blankholder must be used so double-action drawing is required.

When t/D percentage ranges from 0.50 to 1.50

Wrinkling is moderate and low blankholder forces are permitted. Fewer redraws are needed since compressive loads need not be reduced. When t/D percentage Wrinkling is very slight and single-action dies are ranges from 1.50 to permitted if the compressive loads are reduced by 2.5 having larger punches.
When t/D percentage ranges over 2.50 No wrinkling expected so that a blankhold is unnecessary even with high compressive loads.

Deep drawing of rectangular boxes is a special case of deep drawing. As previously discussed in the breakdown of complex parts (Section 2), only the four corners are considered to be deep drawing. For this reason, the corners are the most critical areas. The number of forming operations depends on the h/r ratio (part depth to vertical corner radius) in Figure 3-30. Forming limits also are provided for height to width (Figure 3-31) and height to radius (Figure 3-32) ratios for trimmed or contoured blanks (H-22, T-9). Traditional cup drawing processes assume that failure will occur at the junction of the punch radius and cup wall. For such a non-steady state operation (K-14, M-18, E-5) the draw ratio is independent of the work hardening exponent (Figure 3-33). If the cup drawing conditions can be changed to duplicate steady state operations (failure at the cup wall and die radius junction) the draw ratio becomes highly dependent upon the work hardening exponent n and very large draw ratios can be achieved (K-14, E-5). Tooling methods to accomplish this change include expanding segmented punches and high hydrostatic pressure (E-5, K-25).

3.4 Generating the Stamping Bottom


The stamping bottom is formed by one of two deformation modes biaxial stretching and embossing. Biaxial stretching is obtained when the sheet is clamped around the die opening. This clamping might be positive holddown due to a binder ring or it might be the result of a tight radius preventing metal flow. Deformation in pure stretching is the result of metal thinning; there is no bending component in this deformation mode. Biaxial stretching limits are defined by Forming Limit Diagrams illustrated in Section 4.2. Unfortunately, the distribution of stretch over a punch is very non-uniform and varies form point to point, both in major strain and minor strain. This variation is due to punch geometry, lubrication, and many other factors. The most accurate analysis is accomplished during soft tooling tryout. Some predictive capabilities are available with mathematical modeling programs. A complete hemispherical shape, however, is not achieved for steel except for unusual lubrication conditions. Embossments frequently are used in the bottom of stampings to provide stiffeners, raised mounting of contacting surfaces, clearances for adjacent fixtures, and other uses. Eary (E-2) provides the following design guidelines:
Height of V and Flat V bead = 3t for AKDQ steel

Height of offset

= 2t for CQ steel = 0.8 (R1 + R2) for AKDQ steel


= 0.5 (R1 + R2) for CQ steel

where R1 and R2 are the punch and die radii, respectively. Sometimes the geometry of the bottom is too severe to be generated in a single forming operation. In this case one or more preforms are needed to create the required length of line without a severe strain localization. Once created, the metal can be redistributed into the desired geometry. One of the most common preforms is a dome stretched into the bottom of the stamping. The metal then can be distributed into domes smaller radii or other more complex shapes. This is shown in terms of strain distributions in Figure 3-34. One study documents pressing of ridges into flat panels A-1), while another concentrates on static and dynamic punch stretching of thin diaphragms (G-17). An illustrative example of the use of preforms is provided in a case history of an automotive brake backing plate for drum brakes, Figure 3-34, where breakage was occurring in the seventh die (K-24). The first two forming operations were a domed preform in the first die and a restrike for shape in the second die. The length of line is proportional to the area under the strain distribution curve (average strain times gage length) shown in Figure 3-34. Some area under the curve was created during the preform. However, note the substantial area under the curve required during the second operation or restrike. Because this additional stretch component was created with a sharp radius on the punch, severe strain localization was created. The problem was solved by increasing the depth of the preform to increase the area under the curve after the first die. In addition, the shape of the punch was modified to create additional deformation over the pole (center) of the punch. Modifications were continued until the area under the curve for the first operation was equal to the area required for the second operation. Sometimes special die techniques are required to avoid a severe strain gradient as shown in Figure 3-34. These include local chilling of the punch to increase the strength of the steel at that location (G-32, S-6) and roughening the radius of the punch.

3.5 Interaction of Stretch and Draw

A section taken through the corners of many common automotive stampings would look like Figure 3-35. The geometry is created primarily by the combination of stretch and draw. The two forming modes are interactive. For example, to increase the amount of draw, a larger dimensioned blank is required. To pull this larger blank into the die opening requires a larger force. In transmitting the larger force over the punch radius, additional stretch will be generated sometimes causing failure. An analysis of the design and process variables affecting stretch and draw is shown in Figure 3-36. Examination of the table reveals that these variables most often have an opposite effect on stretch and draw. This is the reason why changes in the design and process variables may not create additional depth of stamping. An illustration is the addition of a domed bottom to a cylindrical drawn cup (Figure 3-37). The draw forming ratio (critical blank diameter or C.B.D. divided by the male punch diameter) is independent of yield strength for a flat bottom cup. When a round bottom is added to the punch, the maximum blank reduction decreases for increased yield strength. This occurs because increased yield strength reduces the work hardening exponent of the steel; this, in turn, reduces the height of the Forming Limit Diagram as well as the strain distributing capacity of the steel. Both of these effects reduce the maximum load capacity of the dome and thereby the maximum blank diameter which can be pulled into the die opening.

3.6 Summary
Existing part and die design limits generally have been created from single forming mode laboratory tests. The number of such configurations and the possible combinations of dimensional parameters actually evaluated in the laboratory has been severely restricted. For example, pure stretch forming laboratory tests generally have been limited to one-inch (25.4mm) and four-inch (101.6mm) hemispherical domes. Tests with compound radii of curvature of different configurations are rarely conducted and reported. The applicability of these test data to production stampings likewise is questionable. For example, only a few test parameters are defined, much less controlled or monitored. Absolute design values, therefore, are difficult to obtain. Studies of the interaction of forming modes necessary to generate complex parts also have been limited. The possible combinations of forming modes, configurations, and dimensions would require a test program of immense proportions, as well as create problems in data analysis and cataloging. For the above reasons, the available part and die design factors or rules of thumb are very limited in their applicability to current press shop designs. Figure 3-1: Nesting irregular blanks to reduce offal.

Figure 3-2: Reduction in blank edge stretchability due to blanking damage(D-18).

Figure 3-3: Measured hold elongation (blanked edge elongation) as a funciton of hole quality and steel grade (H-29).

Figure 3-4: Blanking damage, as evidenced by burr height, can greatly reduce the percent elongation the blanked edge will tolerate in tension before failure (K-24).

Figure 3-5: Edge condition as a function of blanking clearance for various metals (H-40).

Figure 3-6: The load-penetration curves for large and small punch clearances during blanking (H40).

Figure 3-7: Schematic of stamping with differenct geometries.

Figure 3-8: The bending strains on the surfaces increase as the bend angle is increased (S-22)

Figure 3-9: The maxumem tangential strain decreases for a bend angle as the ratio of die radius to sheet thickness (r/t) increases (S-22).

Figure 3-10: Outer fiver strain is shown to be independent of the r/t radio when the curve is normalized for the inverse convex radius of curvature (S-22).

Figure 3-11: After a certain bend depth (or angle) the peak strain no longer increases but the peak width begins to increase (K-24).

Figure 3-12: The maximum strain a bend will achieve is reduced as the ratio of the bend radius to sheet thickness is increased (K-24).

Figure 3-13: As the bend radius increases, the strain distribution becomes very sharp (F-2).

Figure 3-14: Bendability as a function of total elongation and transverse reduction in area for

various strength steels (T-8).

Figure 3-15: Stringer inclusions reduce bendability (T-8).

Figure 3-16: Using length of edge cracking as a measure, sulfide shape control (SSC) provides the greatest improvement at the higher strength levels of steel (K-55).

Figure 3-17: The critical radio of bend radius to sheet thickness, r/t, as a function of total elongation and percent reduction in area is shown for a wide variety of materials (Y-3).

Figure 3-18: Bend-and-straighten operation involves a bending of the sheet to conform to the contour of the die radius and a restraightening of the sheet in the stamping wall.

Figure 3-19: Plan view of a stamping showing the cup draw segments in the corners.

Figure 3-20: The desired punch radius as a function of sheet thickness increases as the sheet

thickness decreases (S-39).

Figure 3-21: The number of redraws to obtain a desired height to diameter ratio (h/d) is given as a function of the flange width to diameter ratio (f/d) (H-34).

Figure 3-22: The height of ears in a deep drawn cup is dependent upon planar anisotropy (W-22).

Figure 3-23: The percent earing increases as the draw ratio increases for aluminum (B-14).

Figure 3-24: Measurements of earing decrease as percent clearance decreases (B-14).

Figure 3-25: The limiting draw ratio (LDR) shown by the vertical arrows is independent of the work hardening exponent, n (F-8).

Figure 3-26: Blankholder force is limited by both wrinkling and punch nose fracture. The

permissible range of blankholder force decreases as the required depth increases (H-34).

Figure 3-27: Forming parameters can change the workable range of the blank holder limits (H-7).

Figure 3-28: The blankholder limits depend greatly on the grade of steel being drawn (H-7).

Figure 3-29: The optimum draw radius depends on the material thickness (S-39).

Figure 3-30: The number of forming operations is a function of the ratio of part depth to vertical corner radius (S-39).

Figure 3-31: Maximum height to width ratio (h/w) for square or rectangular shells drawn in on

operation as a function of corner radius to width ratio (r/w) (H-34).

Figure 3-32: Permissible height to corner radius ratio (h'r) is given as a function of the amount of corner cutting (R) for drawing a rectangular box (H-34).

Figure 3-33: Moving the failure zone from the punch radius to the die radius, the LDR becomes

dependent on the n value and much larter LDR's are possible (K-14, M-18).

Figure 3-34: Strain distributions for the first and second forming operations of an automotive break backing plate (K-24).

Figure 3-35: The corner section of many stampings is delicate balance between stretching over

the punch and drawign to form the wall.

Figure 3-36: Effect of different forming variables on stretch and draw forming modes (E-1).

Within the framework of the sheet metal forming system, the material provides only one component of this system. The material component, however, is the component which is most readily understood and for which the most research has been completed. This is especially true when sheet steel is the material being utilized. This section of the review is subdivided into five parts. They are: 4.1 - TENSILE TEST - The tensile test best defines the parameters which relate to sheet metal formability. The parameters are basic characteristics of the sheet steel and are determined independent of surface interactions with the deforming tools. These parameters are becoming especially important to define the characteristics of the sheet steel relative to mathematical modeling and other forming simulative programs.

4.2 - FORMING LIMIT DIAGRAM - The Forming Limit Diagrams provide a measure of the severity of any sheet metal forming operation. They are the bench mark against which the severity of the forming system is measured. Again, the Forming Limit Diagrams are utilized in most mathematical modeling programs. 4.3 - SIMULATIVE TESTS - Arguments have been made that real world forming involves strain gradients, specimen curvature, biaxial stress states, and the all important frictional interaction of the sheet/tool interface. This section defines a number of the common simulative tests and explores their advantages and disadvantages. 4.4 - COATED STEELS - The coated steels, especially galvanized steels, are being increasingly specified for automotive stampings. This section details one approach being taken to understanding the forming behavior of these materials. 4.5 - HIGHER STRENGTH STEELS - Higher strength steels follow the same rules of sheet metal formability as lower strength steels; the forming parameters simply have different values.

4.1 The Tension Test


4.1.1 INTRODUCTION
The tension test is by far the most popular test used around the world to characterize metal. This is especially true of sheet metal because the thickness of the sheet limits the types of test which can be performed. The tension test is independent of end use; it does not relate directly to forming domes, cups, bends, and the like. Yet the availability of tensile test equipment, the relative ease of performing the test, and the tremendous bulk of test data generated make the tension test the principal candidate for the much sought after single formability test. Unfortunately, the popularity of the tension test also has created a number of problems. The test is not as simple to interpret as it is to perform. Data can be incorrect, misinterpreted, misused, and otherwise abused. A classic story is the excellent one-to-one tension test relationship shown between the anisotropy ratio, rm, and the work hardening exponent, n, for common low strength steels a relationship not apparent from physical metallurgy principles. The problem was traced to one of sampling, for the two steels tested were an aluminum-killed steel (high n, high rm) and an aged rimmed steel (low n, low rm). The problem takes on an even wider dimension for misinterpretation when attempts are made to correlate the tension test with real world forming problems. This section will begin with the basic test procedures, continue with common tensile properties, and conclude with complicating factors such as biaxiality, temperature, test speed, and prior cold work. Since this report is targeted principally for the automotive industry, a mechanical metallurgy approach will be taken, which means that the properties themselves will be taken as the starting point of the review. Almost no discussion will be given to the various chemistry and processing variables which can be modified to create different levels of the properties.

4.1.2 TEST PROCEDURES


The test procedures are well documented by ASTM (A-16), S'E (S-26), and other groups (D-15, S-25, S27) and need not be discussed here. To generate large quantities of less expensive data, a number of companies are using automated tensile machines and computer calculation of data (G-21). This has the important advantage of minimizing operator error or operator variability. Even more important, computerized data acquisition and processing systems have allowed analyses to be conducted that previously were impossible, or at least very difficult and tedious, to perform. One such calculation is the change of the instantaneous work hardening exponent, n, which strain. Another is the

instantaneous plastic anisotropy ratio, r, obtained from a biaxial extensometer. These important measurements will be detailed later in the section.

4.1.3 STRESS-STRAIN CURVES


The primary output from the tension test is a measure of load required to elongate the specimen as a function of the specimen elongation. This information can be presented in a number of ways. The most common is a plot engineering stress against engineering strain (Figure 4.1-1). Engineering stress, E, is defined as load per unit area. More specifically, E = P/A0, where P = instantaneous load and A0 = initial area. Engineering strain, e, is defined as the ratio of increase in length of line to original length. From these curves a number parameters can be measured as shown in the schematics in Figure 4.1-1. One parameter is the yield strength, which is a 0.2 percent offset yield stress (or 0.5 percent offset or other agreed upon number) for steels without yield point elongation (YPE) or is the average stress for steels which have YPE. Other parameters, easily identified in Figure 4.1-1 are ultimate tensile strength ( TS), uniform elongation, total elongation, and fracture. True-stress true-strain curves are less commonly used (Figure 4.1-2a). True stress is defined as the ratio of load to instantaneous area, or T = P/A. Since A decrease faster than P decreases in the necking regime after maximum load on the engineering stress-strain curve, the true stress-true strain curve continues to increase over the entire straining region until the test is terminated by fracture. True strain is defined as the ratio of increase in length to instantaneous length. Thus, true strain, is = ln l/lo. Therefor, tensile true strain is less than tensile engineering strain and compressive true strain is greater than compressive engineering strain. This is based on whether l becomes greater (tensile) or less (compressive) than lo. When log true stress is plotted as a function of log true strain (Figure 4.1-2b), additional information can be made as to whether the plastic deformation can be described by the common parabolic hardening law: T = K n.The values of n and K can be measured. These parameters describe quite accurately the strain hardening behavior of many steels. However, non-parabolic behavior is found for dual-phase steels (B-20, G-22, K-31) and many automotive aluminum alloys (G-10). Two problems occur for non-parabolic hardening metals. First, the condition for the onset of necking is not predicted by the average n value. In this case, the onset of necking must be obtained by graphical solution of the equations: dT/d = T for diffuse necking. dT/d = T /2 for localized necking. This is schematically shown in Figure 4.1-3 (G-10) where dT/d is the instantaneous change of true stress for a change in true strain. The terminal n value usually is lower than the average n value and necking occurs at a lower than the average n value and necking occurs at a lower strain than expected. Second, since the n value is related to the ability of the metal to uniformly distribute strain, changes in n value as a function of strain directly affect its stretch formability. A high initial n value is considered by Rashid (R-5) to be an important factor in the excellent formability of dual-phase steels. Similar data were reported by Bucher (B-20). A whole field of mathematical plasticity has evolved called constitutive equations, which attempt to describe mathematically the metal flow behavior. More than twenty such work hardening laws can be found in the literature (T-15). Four of the most common laws (Hollomon, Ludwik, Voce and Krupskowski/Swift) have been reviewed recently by Ratke and Welch (R-10) and by Truszkowski (T-17).

This field of plasticity also was reviewed at the General Motors Research Symposium on Mechanics of Sheet Metal Forming (K-48), especially the papers by Hutchinson and Neale (H-44, H-45, H-46), Needleman (N-10), and Tozawa (T-13). Additional studies have also been reported by Mohammed (M-33) and Wagoner (W-2). However, detailed review of this specialized area is considered beyond the scope of this report.

4.1.4 COMMON TENSILE PROPERTIES 4.1.4.1 Yield Strength


Much emphasis in North America is placed on yield strength determination because many metals (especially higher strength steels and automotive alloys) are specified by yield strength. However, the yield strength can be difficult to define. For steels without YPE (Figure 4.1-1a) the yield strength is some arbitrary number, such as 0.2 percent offset form the modulus line. Steels with YPE have an upper and lower yield strength. The upper yield strength is so dependent on test conditions that any usefulness of the value is questionable; unfortunately, too many upper yield strength values are measured and reported. While the lower yield strength appears to be an easily measured parameter in Figure 4.1-1b, the value is sometimes hard to define in practice because it changes radically depending on the number of active Lders bands (C-6). A parameter often misunderstood by users of sheet metal is yield point elongation. Special attempts should be made to present the interaction of YPE and formability in an understandable manner. A number of specific areas are: a) The difference between YPE in an as-produced product and the return of YPE as a time-temperature event (B-10). b) Discussion of the presence and absence of YPE as a function of different grades of steel (rim, AK, IF, and dual-phase) and aluminum alloys (5182-0 and 5182-SSF). c) Interrelationship of YPE, Lders bands (see Figure 4.1-4), worms, coil breaks, etc. d) Description of the different methods of removing YPE (temper rolling, flex rolling, tension leveling and others), how they work, speed of return of YPE, etc. (R-18). e) The reduction in the yield strength by removal of YPE (Figure 4.1-5). While this may assist formability by reducing buckle formation, etc., other formability parameters are not restored by temper rolling (C-5). f) The increase in YPE as grain size is reduced is displayed in Figure 4.1-6 (B-10). This is especially important in higher strength steels since they have finer grain sizes. All of the above information is know within the steel industry but has not been definitively accumulated in a single source easily accessible to the users of sheet steel. Rockwell hardness measurements are related directly to the tensile strength of steel. Relationships between Rockwell hardness values and yield strength are less reliable (K-9). However, a special three scale hardness measurement procedure has been published (G-4) which claims to provide a good measure of the yield strength. Experience with this procedure is far too limited in scope and number of tests to comment on its reliability.

New emphasis is being placed on the wrinkle limit of the steel. While fracture is one forming limit, the other forming limit is wrinkling. Wrinkling has been related to the yield point of the steel (F-12).

4.1.4.2 Tensile Strength


In North America there is limited interest in the tensile strength per se, especially with respect to formability. In Japan, however, higher strength steels are specified by tensile strength is synonymous with the onset of diffuse necking and termination of uniform elongation in the remainder of the specimen for metals with parabolic hardening and no strain rate hardening. The tensile strength is a more precise definition and is easier to obtain from test data. Therefore, the popularity of the tensile strength has increased as the prime specification of strength. The ratio of tensile to yield strength (TS/YS) sometimes is used as a measure of stretchability. For metals following parabolic hardening, a theoretical relationship between the YS/YS ratio and the work hardening exponent, n, can be derived provided no yield point elongation is present (Figure 4.1-7). Experimental results show a similar relationship (K-31). A series of curves has been presented for steels with different values of YPE (k-31). Thus, for a given steel with a fixed n value, the TS/YS ratio increases as temper rolling removes YPE.

4.1.4.3 Work Hardening Exponent


The work hardening exponent or n value is important as a parameter that can be related to the ability of sheet metal to distribute strain more uniformly in the presence of a stress gradient (K-24, H-22); a high n value is desirable. The n value has been shown by Backofen (K-29) to depend strongly on the yield strength of steel (Figure 4.1-8). A primary method to increase the strength of steel is to reduce the mean ferrite path, which in turn would reduce the n value (Figure 4.1-9). The availability of computerized data acquisition systems has facilitated a closer examination of formability parameters as a function of strain. The work hardening exponent, n, seems to vary with strain for a wide range of sheet steel products. Thus Lieberg, and Beyer have suggested a new parameter to rank sheet steel stretchability (L-11). Their Om is the logarithmic strain associated with the maximum force in the tensile test; determination of the value does not require an extensometer, a uniform elongation, or adherence to any power law. Examination of the n value for many steels will show that it varies as a function of strain for more than dual-phase steels. Lack of agreement between laboratories in n value determination has been shown to be caused by two-point n value determinations made at different portions of the stress-strain curve (K-31, T-22). Perhaps the n value derived from the initial portion of the curve will correlate with springback and other low strain phenomena. The n value at the onset of localized necking establishes the height of the Forming Limit Diagram. The distribution of strain in the stamping probably is related to some weighted average of the n value over the entire strain history.

4.1.4.4 Uniform Elongation


For metals following the equation T = K n with no strain rate hardening, elongation is related to the n value by the relation n = ln (1 + eu) where eu is the percent uniform elongation divided by 100. More practically, uniform elongation measurements are taken directly from the specimen or the load-elongation chart. However, a very flat load maximum common to sheet steels presents a problem in identifying the exact end point to measure.

Uniform elongation, and stretchability, decrease with increasing yield strength. Typical uniform elongation data are shown in Figure 4.1-10.

4.1.4.5 Localized Necking and Fracture


The formation of the localized neck in a tension test received relatively little interest. The localized neck is extremely important in sheet forming operations, however, and will be discussed later. Deformation continues in the diffuse neck under steadily falling loads. Strain continues to localize with the eventual formation of a thickness neck occurring in the tensile specimen. The dT/d = T /2 relation stated in Section 4.1.3 translates into a true strain equal to twice the value of the work hardening exponent (n) for the onset of a localized neck in power law hardening materials with no strain rate hardening. Deformation in the localized neck continues until fracture terminates all further deformation.

4.1.4.6 Strain Rate Hardening Exponent


There are two primary sources of hardening during deformation. The first is the work hardening previously described by the work hardening exponent (n). A second and parallel source of hardening is strain rate hardening described by the strain rate hardening exponent (m). A simplified constitutive equation is given by: T = K n
m

were K is a proportionality constant and is the strain rate (G-2). Typical values can be n=O and m=0.5 for superplastic material, while n=0.2 and m=0 for most low strength steels at room temperature. Two methods are commonly used to obtain the m value (Figure 4.1-11). One method is to obtain stressstrain curves at different speeds. The difference in yield stress at a given strain can be measured for different strain rate changes (C-8). An argument opposing the use of this method (k-67) is that different microstructures in the different specimens have undergone different strain-strain-rate histories. A method to overcome this (Figure 4.1-11b) is to rapidly change the testing speed during the course of testing a single specimen. This is easily accomplished on a screw driven tensile test machine with a decade speed control or a computer controlled, hydraulic driven tensile test machine. It has been known for some time (S-9, C-8, J-6, B-1, H-24, D-16, G-33) that some common tensile properties change as a function of speed. These are shown for an HSLA steel in Figure 4.1-12 and an aluminum alloy in Figure 4.1-13. The strain rate effects have been extensively evaluated (S-9, C-8). These studies reported that the behavior shown in Figure 4.1-12 was typical of all steel tested. It should be noted, however, that the curve labeled uniform elongation in reality is the elongation at the ultimate tensile strength. The elongation at ultimate tensile strength is the uniform elongation only for metals with a zero strain rate hardening exponent, m. Additional work to determine the effect of cold work and aging of steel (Figure 4.1-14) followed the same pattern. The extensive data are summarized in Figure 4.1-15, which shows the m value (obtained by testing different specimens at different speeds in Figure 4.1-11a) is a function of the static flow stress. Automotive aluminum alloys have m values approaching zero or slightly negative (C-8, G-15). Thus, the data shown in Figure 4.1-13 are typical for these metals. The previous discussion highlighted the influence of strain rate hardening on tensile properties for increasing testing speed. An even more important strain rate hardening effect takes place during a constant cumulative elongation rate test, during which the cross head speed of the tensile test machine remains constant. In the workpiece, the region undergoing thinning strain hardens and becomes resistant to additional deformation. This forces the deformation to the less deformed neighboring elements. In a like

manner, as the local strain rate increases in areas undergoing thinning (or necking), the strain rate hardening in these regions forces deformation to occur in areas undergoing a slow down in the rate of deformation (outside of the neck). For metal s with a positive m value, the neck is turned off or at least delayed. For metals with a negative m value, the growth of the neck is accelerated by strain rate softening. The stress-strain curves in Figure 4.1-16 emphasize the influence of the m value. Even a small m can produce quasistable flow (characterized by a nearly constant maximum load) over an appreciable strain range (G-10). This quasistable flow or post uniform elongation is shown in Figure 4.1-17 as a function of m value. An m value as low as 0.012 can postpone the onset of localized necking long enough to accumulate 40 percent of the specimens total elongation. In a dispersion-hardened zinc (Zn-Ti alloy) with a very low n of 0.05 and an m value of 0.06, nearly 90 percent of the total elongation is post-uniform deformation. In contrast, materials with a negative m value (Figure 4.1-16) rapidly localize deformation beyond maximum load. The strain rate hardening contribution to the Forming Limit Diagram is important (S-4, M-38). This especially applies when comparing FLDs for different classes of metals for which significant differences of m value can occur for equal n values. Muschenborn and Sonne (m-38) even suggest that the formability of heavily temper rolled steel is not reduced as much as the lowered n value suggests. They indicate that temper reductions less than 10 percent do little to affect the diffuse necking component of the deformation and the evaluating stretchability only by the n value may be in error.

4.1.4.7 Total Elongation


The total elongation is widely used as a ductility and formability parameter. The reasoning is simple: the more a metal can elongate in a tension test before fracture, the more it should elongate in a forming operation. As will be seen later, this is an oversimplified approach but one which has merit in several specific cases. The total elongation strongly depends on specimen geometry. While a number of equations have been developed to correct total elongation values for different geometry, the following equation (K-57) appears to make the best overall correction: L12/A1 = L22/A2 where L1 and A1 are gage length and cross-sectional area of geometry 1 and L2 and A2 are gage length and cross-sectional area of geometry 2. Note, however, that other test parameters, such as specimen alignment, strongly affect the total elongation. As the strength of the steel increases, the total elongation decreases (Figure 4.1-10). The amount of reduction in total elongation depends on the strengthening mechanism. Total elongation as a function of yield strength and tensile strength is shown in Figure 4.1-18 and Figure 4.1-19 respectively for various hardening mechanisms.

4.1.4.8 Anisotropy
Most practical metals used by the automotive industry are anisotropic in their properties. This means that properties are different with different directions within the sheet. Two common measures of this behavior are the normal anisotropy ratio rm (resistance to thinning) and the planar anisotropy ratio r. These factors have been well defined and described in terms of test specifications, measurements, and applications (W19, K-24, A-16, W-20, K-20, W-18, S-23).

The r and rm values traditionally have been measured from tensile specimens prepared at 0, 45 , and 90 angles to the rolling direction for steel (K-24, W-19, W-20). Figure 4.1-20 shows that these three directions are excellent representations of the maximum and minimum values for common, low strength 1008 steels. The change of 4 value with angles from the rolling direction is not well defined for other steels; IF steel shows a maximum r value around 55 to the rolling direction (K-31). Other important angles sometimes can exist for aluminum and other FCC metals. The primary method for measurement of anisotropy is the traditional tensile test method. An additional test method is the Modul-r unit. This method may have some advantages for special applications. For magnetic metals, the Modul-r unit is a much more rapid method (M-9). This device measures very accurately Youngs modulus, E, at 0, 45, and 90 degree angles to the rolling direction. From these measurements mean Em and differential E modulus values are calculated. The Em values can be converted to rm by empirical curves, such as shown in Figure 4.1-21. A similar conversion can be made from E tor. Several problems can be encountered with the Modul-r method, however. First, different companies have their own conversion curves in order to normalize new E data to their old manual methods of r determination. Thus, a given set of specimens could be assigned different rm and r values depending on the correction curves used. Second, the scatter band in Figure 4.1-21 is rather wide, rendering the conversion inappropriate for many applications. Third, a fundamental question is raised as to why accurate measurement Em and E should be converted to inaccurate values of rm and r? Perhaps our minds should be recalibrated to think directly in terms of Em and E. In addition, the modulus values for each direction are useful numbers, per se. Finally, the Modul-r unit can not be used with specimen thicknesses greater than 0.08 inches (2mm). This is not a severe restriction, however, because anisotropy values of hot-rolled steels normally have a value approaching one. The Modul-r does solve two problems associated with the traditional tensile specimen method of r value determination. Because the Modul-r unit functions on the oscillating beam theory, no plastic deformation of the specimen occurs. This avoids problems normally associated with cold worked samples, where the uniform elongation is so low that accurate strain measurements are difficult. In addition, the Modul-r test avoids problems often encountered when tensile samples with large amounts of yield point elongation are tested. The traverse of the sample by Luders bands often leads to discontinuities in the width of the sample; these discontinuities can cause significant errors in traditional r value measurements. For hot-rolled steels, only a random orientation of the texture is developed and the rm value is approximately one. This rm = 1 value will persist, independent of yield strength, grain size, etc. (Figure 4.122). Therefore, low strength and high strength steels have equal rm and equal deep drawability. In coldrolled steels, however, the anisotropy is strongly affected by grain size (Figure 4.1-23). Therefore, unlike hot-rolled steels, the r values of cold-rolled steels should be dependent on strength decreasing as the strength of the steel increases. Figure 4.1-24 shows r values for a number of steels as a function of tensile strength. The anisotropy ratios were originally thought to be constant with increasing strain. In fact, one method to obtain accurate r values was to plot width strain as a function of thickness strain for different strain increments; the r value was the slope of the line drawn through the data points. Data (H-39) for steel in Figure 4.1-25 shows a remarkable change in rm for certain steels. This same work generalizes the results as a function of texture strength and documents opposite effects for r values above and below one (Figure 4.1-26). Other researchers (W-15, W-16, L-14) also have noted a change of r value with strain. The argument is proposed by Liu (L-14) that an instantaneous strain ratio should be measured. Grzesik and Vlad (G-41), however, argue that in spite of large changes in the preferred orientation which occurs on straining, relatively little variation in the normal plastic anisotropy takes place. For many steels, the slope of the curve of width strain as a function of thickness strain is constant beyond strain levels of 2 to 3 percent.

However, extrapolating the curve back to the axes does not intersect at the origin. Thus, any curve drawn between the origin and any point on the curve (standard two point method for r value calculation) would create a slope of a different value. Modern biaxial extensometers, electronic data acquisition, and computer analyses allow for such measurements to be easily obtained during a tensile test. Changes in steel compositions and processing can increase the rm values close to the theoretical optimum of 3.0 obtainable in b.c.c. metals (Figure 4.1-27). An r value of 2.8 has been reported (B-14).

4.1.4.9 Shear Strength


An important parameter in calculation of blanking press loads, etc. is the shear strength of the metal, which traditionally has been some fraction of the tensile strength of the metal. Although various tables can be found in the literature (H-41, E-2), an easily remembered approximation (E-1) is: Shear Strength = (YS + TS)/2. Miyauchi (M-30) has developed a simple in-plane shear test to evaluate planar shear deformation caused by differential metal flow in a stamping. Bauer (B-8) describes a torsion test of sheet steel which allows for the calculation of shear stress as a function of shear strain. The primary advantages of the torsion test are attainment of much higher strains than are possible with a tensile test and a principal stress state which corresponds to that found in deep drawing. Miyauchi (M-31) proposes a torsion buckling test.

4.1.5 COMPLICATING FACTORS


As described in Section 4.1.4.8 metals used by the automotive industry do not have properties which are the same in all directions. That is, the metals used are anisotropic. In addition, they respond differently to speed of deformation. Practical forming operations are performed with multiple stresses, different amounts of prestrain, and varying temperatures. In this section, these complicating factors are reviewed.

4.1.5.1 Biaxiality
Most forming is done under biaxial stress states and not uniaxial tension. However, the stress-strain formability parameters obtained by uniaxial tensile test have traditionally been applied to deformation induced under plane strain or balanced biaxial stress states. The developing literature (G-16, L-4, S-34, J11, G-10, R-2). however, is beginning to show that this assumption is not correct. Figure 4.1-28 indicates the balanced biaxial stress-strain curve to be significantly above the uniaxial curve. Ghosh (G-16) reports that the n45 value for aluminum-killed steel to be 0.226 in uniaxial tension and 0.302 in biaxial tension. Preliminary results on automotive aluminum alloys (l-4) show the reverse to be true in that the biaxial curve is lower than the uniaxial curve. Sang and Nishihawa (S-5, R-15, C-4, D-13) have designed various experimental techniques to measure properties under different conditions of biaxiality. Much more work needs to be done in this important area (J-3). In an attempt to better describe the relationship between the biaxial stress level versus uniaxial stress level, Yoshida et al (Y-10) have developed an X factor. The X factor takes into consideration the stress ratio dependence of work hardening, by including functions of r value and r-value-like anisotropy and the dependence of n value upon the mode and amount of deformation (Y-10). This X factor is shown to be related to the r value in Figure 4.1-29. Finally, in order to evaluate combined stress states - as opposed to simple uniaxial tension - Tozawa (T-14, T-13) has adhesively bonded stacked sheets together to compress the metal in the plane of the sheet (Figure 4.1-30). Recent work was completed by Jun and Hosford (J-16), who tested as-received and prestained sheet samples in uniaxial tension, plane strain, and balanced biaxial tension. The prestrain also was conducted

in uniaxial tension, plane strain, and balanced biaxial tension to complete a 3 x 3 matrix of prestrain and test conditions.

4.1.5.2 Temperature
Yet another complicating variable affecting the tensile properties of automotive metals is the temperature of deformation. Temperatures during deformation rarely have been measured. Most often, like the m value, the changes have been thought to be too small to be significant. However, experiments have shown that small changes in tensile strength with temperature (Figure 4.1-31) can produce large differences in formability (G-32). These temperature reductions in the sheet metal are made to locally strengthen the sheet metal in areas of high strain. Even without external modification of the temperature, the importance of thermal notches during deformation is being increasingly recognized. An important paper by Ayres (A-33) showed that tensile elongations for 1008 AK steel tested in air are affected adversely by thermal gradients and beneficially by strain-rate sensitivity at strain rates >10-3 s-1. These two effects appear to be competing influences that partially cancel out under non-isothermal conditions. However, the ductility can be improved by preventing thermal gradients by using isothermal water baths; the total elongation increased from 42 percent to 54 percent. Improved formability in sheet metal stampings can be expected by controlling these thermal gradients. Similar conclusions were reached by Kleemola and co-workers (K-39).

4.1.5.3 Test Speed


The speed of testing in the tensile test affects the resultant properties in many ways. Similarly, the speed of forming affects formability in a variety of ways. Test speed is not to be confused with deformation speed. Test speed is global like the rotation of the earth. However, deformation speed can be different for each element of the test specimen. Strain rate has a dual role. One is the global effect which is established by the test speed of the specimen; this will be discussed here. The other is the strain rate effect in the incipient neck and the resulting beneficial effect of strain rate hardening (discussed in Section 4.1.4.6). Increased test speed will increase the yield and tensile strength as shown previously in Figure 4.1-12. Therefore, a standard test speed is important for reproducibility within a laboratory and between laboratories. The ASTM specifies test procedures for this purpose (A-16). However, the question is continually raised whether this speed is indicative of the properties encountered during either the forming cycle or in-service environments such as during crash management (K-52). A good case is made for testing at speeds equivalent to forming speeds, especially for predicting press and tool loading. A new problem now introduced is the ability of performing high speed tensile tests without introducing excessive vibrations into the dynamic tensile test machine (S-8). Other test speed effects are more difficult to measure. Faster test speeds decrease the time for dissipation of heat generated by the deformation. This, in turn, increases the thermal gradients developed along the length of the specimen (A-33). Thermal gradients not only affect the strength of the steel at each point along the specimen, but may achieve sufficiently high temperatures to initiate metallurgical modifications within the specimen.

4.1.5.4 Prior Cold Work


Many metals are deformed after being previously cold worked. A number of studies (B-14, H-20, H-23, T5) have been conducted to evaluate the effect of prior cold work on various properties. Figure 4.1-32 shows the instantaneous n value as a function of strain for a various prior processing studies. Strain rate data for cold work samples are presented in Figure 4.1-14. The drop in uniform strain in Figure 4.1-33 indicates a rapid loss of flow stability when uniaxial tests are conducted after a biaxial prestrain. The residual uniform tensile elongation drops to zero for an effective biaxial prestrain of 0.08. Many questions are unanswered about the effect of prior cold work on metal properties (F-6).

4.1.6 SUMMARY
Of all the common laboratory tests, the tensile test probably is the most defined and best understood test available. The tensile test has certain advantages compared to other laboratory tests, including singular loading mode, no specimen curvature, no interface contact with a forming tool, and a simplified analysis. Numerous properties can be obtained from the tensile test which, in turn, can be correlated to different forming modes in production stampings. Some properties, however, depend on the specific work hardening law used to best describe the actual behavior of the specimen. Once a specific property from the tensile test can be correlated to a specific stamping (such as the normal plastic anisotropy ratio with deep drawability), the problem then becomes one of determining the range of the property (both maximum and minimum) which will assist in generating a satisfactory stamping. Figure 4.1-1 Typical engineering stress-strain curves for steels without and with yield point elongation.

Figure 4.1-2 Comparison of true stress-strain curve with and engineering stress-strain curve.

Figure 4.1-3 The determination of the onset of diffuse necking for metals. The influence of srain rate hardening (m) is shown (G-10).

Figure 4.1-4 The relationship between Lder's bands for steels having yield point elongation (B-

10).

Figure 4.1-5 The lower yield stress decreases as percent temper rolling increases until the yield point elongation is eliminated. Then work hardening causes the yield stress to increase (C-6).

Figure 4.1-6 The relation between yield point elongation and grain size (B-10).

Figure 4.1-7 A theoretical relationship exists between the n value and the TS/YS ratio only for steels with no yield point elongation and parabolic hardening.

Figure 4.1-8 Al linear relationship exists between yield strength and n value for steels with a yield

strength greater than 50 ksi (345 MPa) (K-29).

Figure 4.1-9 The strain hardening exponent n correlates with the log of the mean free ferrite path (B-10).

Figure 4.1.10 Both uniform elongation and total elongation decrease as a function of yield strength (F-2).

Figure 4.1-11 Two methods are commonly used to determine the strain rate hardening exponent, m.

Figure 4.1.12 The yield and tensile sterngths increase as fucntion of test speed for a high strength steel due to a positive m value (C-8).

Figure 4.1-13 The yield and tensile strengths typically remain unchanged with test speed for most

automotive aluminum alloys with no strain rate hardening (C-8).

Figure 4.1-14 Yield strength as a function of strain rate for a cold-rolled rimmed steel sheet (S-9).

Figure 4.1-15 The strain rate hardening exponent, m, can be correlated to the static flow stress (S-

9).

Figure 4.1-16 The shape of the stress-strain curve beyond maximum load (indicated by the vertical arrow is related to the strain rate hardening exponent (m) (G-10).

Figure 4.1-17 Post-uniform elongation is directly related to the strain rate hardening exponent (m) (G-10).

Figure 4.1-18 The total elongation for a given yield strength depends on the strengthening mechanism (K-55).

Figure 4.1-19 Total elongation for a given tensile strength is a function of the hardening

mechanism (M-14).

Figure 4.1-20 The r value for sheet steel depends on specimen angle relative to the rolling direction (F-8).

Figure 4.1-21 The traditional correlation curve between Em from the Module-r unit and manually

determing rm values from tensile specimens (M-35).

Figure 4.1-22 The r values are independent of yield strength for hot-rolled steels (F-2).

Figure 4.1-23 The rm value is a function of grain size for cold-rolled steels (B-10).

Figure 4.1-24 The rm can vary greatly as a function of tensile strength depending on the hardening mechanism (M-14).

Figure 4.1-25 The mean anisotropy ratio, rm, has been shown to be a function of the strain level (H-

39).

Figure 4.1-26 Experimental results show the change of r value as a function of strain (H-39).

Figure 4.1-27 The relative values of r0, r45, and r90 depend upon the chemistry of the steel (B-14).

Figure 4.1-28 For steel, the stress-strain curve for balanced biaxial strain is higher than the curve for uniaxial strain. This causes a strain path change from biaxial prestrain to uniaxial tension to induce a lowering of the FLD (G-16).

Figure 4.1-29 The X factor, the stress ratio dependence of work hardening, is shown as a function

of the rm value (Y-10).

Figure 4.1-30 Sheets of metal are glued together to form pieces thick enough for compressive edge loading (T-14).

Figure 4.1-31 The tensile strength of steel decreases for steel as the test temperature increases

(G-32).

Figure 4.1-32 Variation in n value with strain is shown for different amounts of prestrain temperrolling (B-14).

Figure 4.1-33 The change in properties is given as a function of prestrain. Note how fast the

uniform elongation is reduced to zero (G-16).

4.2 Forming Limit Diagrams


4.2.1 INTRODUCTION
Forming Limit Diagrams (FLDs) have been empirically constructed to describe the strain states, or combinations of major (e1) and minor (e2) principal strains, at which a highly localized zone of thinning, or necking, becomes visible in the surface of sheet metal. The first published FLD for sheet steel appeared in 1963 for stretch forming the low carbon, low strength steel commonly used by the automotive and appliance industry 0.036 inch (0.9mm) 1008 steel (K-15). Derived from laboratory rigid punch dome tests, the first FLD (Figure 4.2-1) showed the maximum strain (called fracture strain) that this steel and other metals could withstand as a function of the principal strain ratio. The criterion used to define the maximum strain was the onset of a band of highly localized thinning in the sheet surface. This criterion was selected because the appearance of a localized thinning or visible neck is sufficient cause for rejection of exposed stampings in the press shop. For unexposed stampings, the formation of the localized thinning signals the termination of further general deformation throughout the stamping and introduces potential structural defects which can affect in-service performance. The empirically derived laboratory data were utilized in production press shop experiments from 19631965. The results of such work were summarized in a 1965 paper which laid the foundation for the present analysis technique (K-12). This paper showed that the maximum strain limits obtained from stretch formed production stampings were identical to those previously obtained in the laboratory. In addition, the ratio of the principal strains plotted on the absicca was replaced by the minor strain. Additional press shop experiments conducted during the 1965 to 1968 period led to two companion papers in 1968. One by Keeler (K-21) detailed research conducted on the "right side of the FLD" where e2>0 for stretch forming and on by Goodwin (G-23) detailed research conducted on the "left side of the FLD" where e2<0 for deep drawing. Taken collectively, the two limiting strain curves represent the onset of localized necking for different values of the minor strain encountered in sheet metal forming. The FLD just described is known as the necking or instability FLD. Care must be taken not to confuse the instability FLDs appearing in the literature with two other FLDs which have appeared in the literature. One is the fracture FLD, measured when the specimen physically separates (G-39, K-35, B-6, N-6, T-3, D-10). The other published FLD is the wrinkling or buckling FLD (H-8, H-33), which details strain states

for which wrinkling is possible. These other two types of FLDs will be described in Sections 4.2-6 and 4.27).

4.2.2 SHAPE OF THE FLD


The Forming Limit Diagram (or Forming Limit Curve in some areas of the world) can be plotted in a variety of ways. The most common is shown in Figure 4.2-2. Here the vertical axis is the largest engineering strain in the plane of the sheet or major strain e1. If the circle is used as the grid for geometry, the major strain is obtained form the long axis of the resulting ellipse. The horizontal axis is the smallest engineering strain in the plane of the sheet or the minor strain e2. This plot encompasses all possible combinations of major and minor strains encountered in practical sheet steel forming; it is sometimes denoted as strain space in the literature. Data to be plotted in the strain space can be obtained in several ways. The most common measurements are made on a series of rectangular strips stretched over a punch. This method was proposed by Nakazima (N-6) and used by Veerman (V-4). Hecker (H-19) developed, streamlined, and studied in depth a comparable method now used by many laboratories. Instead of many punch configurations and test conditions, a single hemispherical punch usually four inches or 100mm is used to stretch to necking gridded sheet specimens securely clamped at the periphery. Different minor strains (e2), and therefore different amounts of strain biaxiality, are obtained by varying the width of the specimen (Figure 4.2-2). Sometimes notched samples are required to avoid die radius failures in narrow strips (A-2). Recently the literature has detailed other test techniques for obtaining onset of instability data. Some investigators use hydraulic bulge tests using masks of elliptical shape with different aspect ratios to obtain a series of positive minor strain values for establishing the right side of the FLD (M-16). Tensile specimens of varying widths with side incisions are used to generate varying degrees of negative minor strain to establish the left side of the FLD (M-16, H-6). A completely different approach is suggested by Gronostajski and Dolny (G-35). They use modified Marciniak domes for their tests. Both the specimens and the spacers have circular arcs of varying radii cut into the width dimension of the sample. This specimen configuration is used to determine both the instability FLD and the fracture FLD. These authors argue that the modified Marciniak domes eliminate friction between material and tool surface, retain the flat surface of the specimens during the whole straining process, closely approach proportional straining for all strain paths, and require only one punch and one die to obtain the full FLD. Similar claims could be made by Azrin (A-35) and Ghosh (G-8, G-12). Their in-plane stretching technique consists of an elliptical groove machined in the center of a sheet which is subsequently stretched over a flat punch. The greatest controversy in FLD determination is the procedure for determining the onset of the instability for which measurements are then plotted to form the FLD. Keeler (K-21) and Hecker (H-19) suggest stopping the deformation at the onset of necking. The strain for the circle directly over the location of the incipient neck is measured and is coded as to unnecked, incipient neck, and necked. The FLD is drawn between the unnecked and necked coded data points and through any incipient neck data points. A similar data acquisition system is used for measurements from production stampings. This technique, though harder to conduct, avoids the problems generated when attempting to measure the next whole circle adjacent to the failure location. The problem with the adjacent circle technique is that it delicately balances the necking strain for circles too close to the neck against the reduced strain level due to gradients for circles too far removed from the neck. Similar problems have been encountered when extrapolation of gradients across the fracture itself have been attempted (I-3). Some strain gradient estimation techniques (A-21) lead to FLDs which are quite different from those obtained previously by classical methods. Keemola and co-workers argue that the change of the strain path towards the plane strain provides the lowest limit strains (K-40). Thus, the definition of the FLD and the measurement technique are critical in determining the shape and level of the FLD.

The steel, the shape of the FLD for proportional straining (linear strain path) shown in Figure 4.2-3 is used as a standard shape in many press shop applications (K-28, K-30), especially in North America. The vertical position of the FLD changes depending on the characteristics of the sheet steel which the FLD represents. In other applications, the FLD for a specific steel sample is empirically obtained in the laboratory through the use of Nakazima samples stretched over rigid punches (H-19). In this case, a regression analysis is used to determine the best fit curve. European determination of the FLD is based more on best-fit curves than on standard shape. In addition, the European curves consistently tend to have the minimum of the FLD at a minor strain level of + 0.05 (J-3, K-40). The cause of this is unclear but may be due to the FLD determination procedures used in Europe. The use of the standard shape curve has simplified press shop application, since the entire curve is now defined by the plane strain intercept (FLD0). However, this standard shape does not apply to stainless steel and other steel alloys (R-6, K-28). A variety of FLD shapes are found for aluminum alloys. Interestingly, however, a number of the automotive aluminum alloys (2036-T4, 6009, 6010, 5182-0, etc.) have FLDs with shapes similar to that of low carbon steel (H-14, S-35, H-16). Another method of portraying the data is to plot the curve on a true major strain true minor strain curve (Figure 4.2-4). The left hand side of the curve has been shown to be a straight line with a slope of -1 when plotted on a major true strain versus minor true strain graph (I-3). This means the left hand curve is equal to the strain allowed in plane strain plus a shear component 1 = 2). The right hand curve remains to be the original curve replotted in the true strain format (K-15, K-12, K-24). The major minor strains in the surface of the sheet metal at the onset of instability can be used to calculate the thickness strain (through the constancy of volume rule). Therefore, the curve in Figure 4.2-3 can be used to generate a thickness strain Forming Limit Diagram as shown in Figure 4.2-5 (K-11). This thickness FLD can be presented on either an engineering or true strain axis. Finally, the Japanese literature portrays the FLD with a rotation of the axis (Figure 4.2-6). Here the traditional North American FLD is mirror imaged (rotated 180 degrees) about the y axis and then rotate 90 degrees clockwise. A review paper on Forming Limit Diagram Sheets was prepared for the 1974 Sagamore Conference (K-28). All aspects of FLDs and the literature as of 1974 are detailed in that document. Therefore, the present review will provide only an update of the literature since the Sagamore Conference, and complements a recent review published by Mellor (M-20).

4.2.3 LEVEL OF THE FLD


The FLD is a characteristic of the steel and is independent of the stamping for which the steel is used. For example, different stampings will have different forming modes which are utilized in the formation of the stamping. These forming modes can be plotted on the Forming Limit Diagrams Figure 4.2-7). Therefore, the higher level of the FLD, the higher the strain which can be imparted to the stamping before failure. The level of the FLD depends on a number of parameters. These are reviewed below.

4.2.3.1 Sheet Thickness


A previous publication (K-28) indicated a sheet thickness correction in which the plane strain intercept (FLD0) increases from a value of n for zero thickness sheet as shown in Figure 4.2-8. Additional studies showing this effect include H-1, C-3, H-32, H-25, K-40. Work on 2036-T4 aluminum (K-31) also has shown an increase in the FLD0 with sheet thickness from an initial value approximately equal to the n value. However, the slope is substantially less. It is suggested that the aluminum and steel both undergo an increase in FLD0 with sheet thickness simply because the neck is geometrically more diffused as the sheet thickness increases. The steel, however, has a large additional thickness influence which this author suggests is related to the strain rate hardening exponent, m.

The role of strain rate sensitivity in the thickness effect of the FLD has been mathematically modeled by Rao and Caturvedi (R-4, R-3). Studies of the thickness effect for dual-phase steels would be important for additional understanding of the physical basis for the thickness effects.

4.2.3.2 Mechanical Properties


A paper at Microalloying-75 (K-29) first indicated a systematic decrease in the FLD0 as the n value of HSLA steels decreased below the 0.21 level. Additional work reported in the discussion of the Microalloying 75 paper (K-29) showed a whole family of sheet thickness n value curves which could be used to determine an FLD0 for any HSLA steel (Figure 4.2-9). Still to be fully explained is the reason why 70-30 brass (Figure 4.2-10) has an FLD almost equal to that of steel, even though its terminal work hardening exponent (n) is almost twice that of AK steel. Some of the problem may be explained by the m value and sheet thickness effects described above but it is uncertain whether that is the total solution (A35, G-12). It might be expected that the normal plastic strain ratio, rm, affects the level of the FLD since the r value is known to be important in tensile tests, deep drawing, and other deformation modes. However, studies have shown the level of the FLD to be independent of the r value (K-28). The r value, however, does affect the strain path taken during deformation. Figure 4.2-11 shows this effect on the strain path taken by two tensile specimens. Thus, although the level of the FLD is independent of the r value, the permissible FLD ceiling strain is indirectly a function of the r value through the strain path (B-18, K-45). The traditional FLD details the strain conditions for the onset of localized necking. Therefore, correlations between the level of the FLD and the uniform elongation are not expected. The uniform elongation does not incorporate the effect of the strain rate sensitivity which tends to delay the onset of the localized neck and increases the increment of post-uniform elongation. This is especially evident when the FLDs of steel and 2036-T4 aluminum are compared (H-19). Both have approximately the same uniform elongation values (work hardening or n value controlled) but widely differing strain rate hardening (m value) characteristics. Likewise, no correlation exists between the total elongation and the traditional FLD based on instability. The total elongation depends on the strain required for the localized neck to progress to fracture. The total elongation decreases with increasing biaxiality, but the FLD first decreases and then increases. The literature recognizes this by publishing fracture FLDs. These are discussed in Section 4.2.6. The literature contains numerous papers and reviews which attempt to show the role, or absence thereof, which steel cleanliness plays on the level and shape of the FLD (K-40, S-16, D-12, A-25, A-11, K-28, S31, T-7, J-2, K-38). No clear, definitive conclusion can be reached.

4.2.3.3 Bending
Bending adds a positive strain component to the convex surface of the sheet and a negative strain component to the concave surface of the sheet. Somewhere between these two extremes is a neutral axis. The compressive (negative) strain and the zero strain in the neutral axis counteract the tensile strain in the convex surface and do not allow a through-thickness localized neck to develop. Strain readings taken on the convex surface (the surface usually selected for strain measurements) can exceed the FLD without the initiation of localized necking. This leads the press shop to consider the FLD to be too conservative (B-12, A-22). Three methods have been used to correct the strain readings obtained on stampings before plotting on the FLD (K-31). The first is to average the strains from the convex and concave surfaces. In this manner the bending strains are subtracted from the convex surface strains, which are then plotted on the FLD. This method does not make any assumptions about the location of the neutral axis.

The second method is to calculate the bending strain from the measurement of the inside radius, the sheet thickness, and an estimation of the location of the neutral axis. Such a calculation will show that a 1t bend will add a strain of 33 percent to the convex surface. The third method is to measure the thickness strain of the sheet, either by ultrasonic thickness measurements, cross-sectioning, or some other method. The major strain component then can be calculated from the thickness strain and the minor strain through the constancy of volume formula.

4.2.4 DEFORMATION PATH AND PRESTRAIN


The initial FLDs were generated by single path deformation modes. These single paths also are known as proportional straining and will plot as a straight line on the FLD. Even here differences exist as to the definition of proportional straining whether the strain path is a straight line in engineering strain or true strain space (R-9). Most experimental research on FLDs in the last decade has been concentrated on the effect of changes in strain path especially radical changes. The two extreme boundary cases for steel (Figure 4.1-12) appear to be the upper curve formed by unixial tension followed by balanced biaxial tension and the lower curve formed by balanced biaxial tension followed by uniaxial tension. These multiple path FLDs have been examined by numerous investigators (K-35, G-16, I-7, K-56, R-7, M-15, M-37, K-37, L-4, K-40, A20, G-37, G-26, G-36). A rule of thumb has been derived which summarized the results of the strain path research (R-9). This rule states that the addition of two linear strain path increments will cause: - higher limit strains if the strain increment ratio is greater in the second stage than in the first stage (clockwise shift in strain path). - lower limit strains if the strain increment ratio of the second stage is less than that of the first stage (counterclockwise shift in the strain path). Matsuoka and Sudo (M-15) provided an interesting suggestion that strain space be divided into three zones. All strains in Zone 1 can be reached safely under all circumstances. Zone 2 contains strain states that can or can not be reached safely depending on strain path. Zone 3 contains strain states that can not be reached under any conditions. In Figure 4.2-12, Zone 1 is below the minimum forming limits curve ( = 0) and Zone 3 is above the maximum forming limits curve ( = 1.0). The problem remains, however, to provide press shop documentation and case histories of the practical application of strain path changes in single or multiple forming stage operations (K-41, K-42, K-56, G-39). Most problem areas in forming can be attributed to deformation in the first major forming die.

4.2.5 MATHEMATICAL DESCRIPTION


A mathematical description and derivation of the FLDs has been the other important research area relative to FLDs during the last decade. A mathematical description of FLDs is important for simplicity of press shop calculations and for entering the FLDs into computer plotting programs for computer graphic output. Here the conversion of the graphical FLD into an equation form is required. This was done initially for an S'E Recommended Practice (S-27). For low strength steels the following equations can be used to approximate the FLD (S-27, K-32): When e2 is 0 to plus 30, FLD = (0.6e2 + 25 + 350t) for t in inches

FLD = (0.6e2 + 25 + 13.8t) for t in mm. When e2 is 0 to minus 30, FLD = (1.5e2 + 25 + 350t) for t in inches FLD = (1.5e2 + 35 + 13.8t) for t in mm. Note that major and minor strains are expressed as a percent (not a decimal value) and the sign of e2 is disregarded. A more fundamental approach is the mathematical prediction of FLDs from basic plasticity analyses of instability. A number of papers have addressed this problem (M-7, M-8, J-3, K-40, L-1, G-20, V-5, T-5) and it was reviewed at the General Motors Symposium on Mechanics of Steel Metal Formability (K-48). In general, problems have occurred when experimental FLDs are compared to mathematical models derived from theory. While numerous factors have been introduced into the equations, the physical significance of many of these factors eludes many readers (K-40, B-19, D-10). A solid, physical understanding of the FLD still is lacking.

4.2.6 FRACTURE FORMING LIMIT DIAGRAMS


The previous discussion has been restricted to FLDs derived for onset of a localized neck. The literature also contains FLDs derived for the actual fracture of the specimen (P-3, B-6, N-6, T-3, D-10, I-3, G-39, K35). Fracture strain measurements are very sensitive to a number of test parameters. Grumbach and Sanz (G39) showed the grid length used for measurements not only influenced the level of the FLD-fracture but also the shape (Figure 4.1-13). For zero gage length the fracture strain (e1) decreased uniformly as the degree of biaxiality is increased. As the gage length is increased, the FLD-fracture measurements included an increasing proportion of the uniform and post-uniform strain. When the gage length becomes large enough, the FLD-fracture duplicated the FLD-localized necking. Baret and Wybo (B-6) observed a grid length effect on the FLD-fracture but no FLD-localized necking. Other factors which influence the FLD-fracture are steel grade, inclusion size and shape, sample orientation, degree of biaxiality, etc. (N-6, G-39, K-35, D-10, F-1). Because of this extreme sensitivity, it is conceivable that an FLD-fracture would have to be obtained not only for every coil of metal but for locations along the length of the coil to be truly representative of the coil. The fracture FLD, therefore, seldom is utilized in the press shop. First, the fracture FLD will vary from point to point in a coil of steel as a function of cleanliness. This makes it difficult to generate a FLD representative of the coil. Second, necking is cause of rejection in many stampings, especially for those exposed applications. Third, once the localized neck has begun, all deformation is now confined within the neck. Therefore, even though the strain level at fracture in the neck is much higher than the strain at the onset of the localized neck, the active gage length for the additional straining is so small that little additional stamping depth is achieved. For low strength steel, the fracture FLD is well above the instability FLD (Figure 4.2-14). As expected, the curve decreases continuously with increasing biaxiality (increasing minor strain, e2). For some high strength steels, the fracture FLD has dropped below the localized necking FLD (Figure 4.2-15). This is seen in the laboratory as fracture without a localized neck-typically for large positive and large negative minor strains (K-31). In one special case, the normal sequence of uniform deformation and diffuse necking was interrupted by ductile fracture prior to obtaining strain levels predicted by the FLD-necking (K-31). The steel was a relatively dirty 50 ksi (345 MPa) yield strength steel without inclusion shape control. Major strains of 65 percent were achieved in the rolling direction of the sample with a visible neck beginning to form perpendicular to the major strain direction. Suddenly, a sharp fracture occurred parallel to the rolling

direction and cut through the neck in a perpendicular direction. Strains across the neck were only 40 percent with no evidence of a neck preceding fracture. Based on the above discussion, steel producers should use techniques, especially inclusion shape control, to suppress ductile fracture initiation sufficiently to insure that the FLD-fracture is always at a strain level higher than the FLD-localized necking. This will allow for the maximum stretchability the steel is capable of producing.

4.2.7 BUCKLING/WRINKLING DIAGRAMS


Arguments have been made that fracture is not the only failure in forming. Wrinkles and buckles also contribute to stamping rejection (h-8, H-33). The criterion for the initiation of buckles is that the thickness strain is positive. An increase in thickness means a propensity of the sheet metal to buckling. The strain states which cause an increase in thickness can be shown on the same strain space used for the Forming Limit Diagram (Figure 4.2-16).

4.2.8 SUMMARY
The Forming Limit Diagram represents the combinations of major and minor surface strains which sheet metal can undergo without satisfying the conditions for the onset of localized necking. For this reason the FLD is a formability characteristic of the metal independent of the specific application of the metal. The data points plotted on the FLD are values obtained from the actual stamping under investigation. The proximity of the various data points to the failure line indicated the severity of the stamping. For steel, a standard shaped FLD curve is used to make the analysis technique simple and useful for press shop application. The height of the curve-specified by FLD0 is determined from a nomograph based on the thickness and the work hardening exponent of the steel. These quick analysis techniques are not applicable to alloys of steel or other metals. For these metals, experimental FLDs must be obtained in the laboratory. Additional Forming Limit Diagram research is still required for more complex forming conditions, such a prestrained metal and radical changes in strain path. Figure 4.2-1 First published Forming Limit Diagram showing major strain as a function of the

strain ratio (K-15).

Figure 4.2-2 A Forming Limit Diagram commonly is generated in the laboratory by clamping and stretching strips of varying width. The strain at the onset of localized necking (failure points in the

diagram) is poltted as a function of the minor strain (H-19).

Figure 4.2-3 A standard shaped Forming Limit Deagram is used in many North American press shops. The height of the curve is set by the FLD0 determined from the nomograph in Figure 4.2-9

(A-12).

Figure 4.2-4 The axes of the traditional Forming Limit Diagram are engineering strains. When plotted on true strain axes, the left side of the standard shaped FLD becomes a straight line at 45 degrees to the vertical axis (-34).

Figure 4.2-5 The conditions for the onset of localized necking (FLD) can be plotted in terms of the thickness ratio and the minor strain. The conversion is accomplished with the constancy of volume formula. The left side of the standard shaped FLD now becomes a constant thickness

ratio (K-11).

Figure 4.2-6 The Japanese version of the FLD is plotted with a different orientation of the axes (I7).

Figure 4.2-7 Different forming modes can be depicted as different strain paths on the FLD.

Figure 4.2-8 Effects of sheet thickness on the FLD0 for steel and aluminum (K-31).

Figure 4.2-9 A nomograph to obtain the FLD0 for a steel when the values of the work hardening

exponent, n, and the sheet thickness are known (K-29).

Figure 4.2-10 The level of the FLD's for different metals depends upon both the work hardening exponent, n, and the strain rate hardening exponent, m (G-10).

Figure 4.2-11 The plastic anisotropy ratio, r, determines the strain path a tensile test and other

deformation modes will follow prior to the onset of localized necking.

Figure 4.2-12 A variety of "FLD's" are possible if a radical change in strain path takes place (K-56).

Figure 4.2-13 True fracture strain as a function of the minor strain and the gage length used for the

strain measurements (G-39).

Figure 4.2-14 Both the FLD-fracture and the FLD-localized necking are shown for an AK DQ steel (G-14).

Figure 4.2-15 A high strength steel with a large number of inclusions can have a FLD-fracture

which will intersect the FLD-localized necking (K-31).

Figure 4.2-16 The "wrinkling FLD" delineates conditions for which buckles or wrinkles can occur. These conditions generally are those strain states which cause an increase in thickness.

4.3 Simulative Tests


4.3.1 INTRODUCTION
The Forming System is a complex interaction of material, lubricant, die, and press. By holding the incoming material constant, one of the primary components of the Forming System is held constant. This, in theory, should reduce one of the causes of press shop breakage. Therefore, press shops continually search for a test(s) which will permit evaluation of the formability of the incoming material. Formability of incoming steel can be evaluated by three general types of tests. These are: 1) fundamental properties, 2) simulative tests, and 3) actual press shop trials.

Fundamental properties utilized for formability evaluation generally are those properties which have previously been shown to correlate to formability. Primary properties include work hardening exponent (n), strain rate hardening exponent (m), and anisotropy ratio (r). Other related properties are yield strength, uniform elongation, and total elongation. These properties are measured from a tensile specimen, which is deformed at a constant rate by a uniaxial stress in free space (no surface contact). These properties are widely used to assess the formability of sheet metal. Two major determinations must now be made. The first is choosing the correct property or properties which can predict the performance of the sheet metal in stamping under investigation. For example, the work hardening exponent, n, is related to stretch forming but not to cup drawing. The problem now becomes one of determining whether the critical zone in the stamping is formed by the stretch forming mode. This may be difficult without an extensive evaluation of both the stress and strain states. The identical strain state can be created by either a stretch forming mode or a compressive mode. Tensile instability terminates the former but not the latter. The second determination is selection of the proper test parameters. An argument is made that most sheet metal is formed with a biaxial stress state, while the traditional tensile test are uniaxial. A valid question therefore is whether various degrees of biaxiality change the n value. If so, a multiplicity of n value test would be required to match the exact biaxial stretch conditions of the stamping for which the steel is intended. Other test parameters must also be selected for effective correlation with the stamping. These include deformation speed, test temperatures, and prior deformation conditions. The argument also is made that the surface/lubrication interactions are important in forming of actual stampings and these interactions are absent in fundamental property tests such as the tensile test. To counter this last complaint, some measurements of surface morphology and coefficients of friction can be made for each sheet material being characterized. On the other extreme, press shop trials indicate the response of the total Forming System. However, failure of the system to produce a satisfactory stamping does not provide information as to which component(s) is at fault. The Forming System is, unfortunately, a dynamic system. Failure of the system to produce a satisfactory stamping simply means that the system components are mismatched at the time the stamping is made. Without extensive system checks it is difficult to say which component such as the material is at fault. Even worse, the current make/break evaluation method does not provide sufficient gradation between materials with different levels of formability. Another problem with press shop trials is the need to create a full size blank and bring it to the press. Finally, interjecting press shop trials into a production schedule is sometimes difficult. Simulative formability tests were designed to overcome these problems a compromise between fundamental property evaluation and actual and actual press trials.

4.3.2 USES OF SIMULATIVE TESTS


Simulative tests can be used for three purposes in sheet metal forming: a. Evaluate incoming metal with respect to a historical baseline. b. Predict the success of the metal in a specific press application. c. Analyze the effects of modifying various parameters.
a) Evaluation of incoming metal with respect to a historical baseline may be a simple go/no-go test or a more sophisticated Statistical Process Control charting. In this application, the simulative test indicates whether the incoming steel has the same characteristic values as the previous shipments of steel. The simulative test used in the screening may have little direct relationship to the actual forming operation. For example, a steel may be subjected to the Limiting Draw Ratio test as determined by the cylindrical cup draw or Swift flat bottom cup test. The critical mode in the actual production stamping may be breakage along a blanked edge, which

depends on the work hardening exponent, n, and the normal plastic anisotropy ratio, rm. The rm also is a controlling factor in the cup drawing operation. Therefore, the cup test results can be indirectly related to the blanked edge tearing. More importantly, historical records for the stamping in question may show that steel from Company A processed to give the necessary minimum rm value also will have a sufficiently high n value to make the stamping.

In the above example, the Limiting Draw Ratio evaluation was used to check whether the incoming steel probably was processed identically to the previous shipments of steel. The required minimum rm value obtained on steel from another company may not insure successful production of the stamping in question. Different processing cycles may create identical rm values with entirely different work hardening exponents. For evaluating incoming metal with respect to its historical baseline, the application of simulative test results is very narrow in scope and requires extensive historical records. In addition, no statements can be made as to whether steel outside the limits will or will not make the stamping successfully. b) Prediction of successful forming by a given steel in a specific stamping is very tenuous when made by the simulative test route. The relationship between the simulative test and the actual stamping usually is marginal. For example, stretching over a one-inch (25mm) diameter punch may not be related to a stamping in which the largest radius is twice metal thickness or 0.04 inch (1mm). Likewise, a specific lubricant may assist forming over a generous radius but may restrict forming over a tight radius. c) Analyzing the effect of various forming parameters may be the most effective utilization of simulative tests. Here the different deformation modes can be isolated and the effect of material and process variables can be studied. The effects of material properties, lubricants, punch speed, die radii, temperature, and all other Forming System variables can be studied without interaction from other forming modes. The results then be extrapolated to the more complex forming processes. The study of forming parameters with simulative tests also provides verification data for the various mathematical models of forming. The simulative tests usually are axisymmetrical, which is an easier form to model. The actual dimensions of the simulative test can be entered into the mathematical mode. The strain histories predicted from the model can be verified with actual strain readings taken from grids on deformed samples.
The frictional conditions at the interface between the workpiece and the tool are the least quantified of any of the forming variables. As such, the correlation between simulative tests and production stampings depends on how well the simulative test reproduces the interface friction. The same comments apply to the mathematical models.

4.3.3 FORMING MODES


The most important description of any simulative test is the forming mode which it simulates. Some tests are single mode tests, while others are combinations of different modes. SINGLE MODE STRETCH

Hole Expansion, Yoshida Buckling (minor strain is negative), LDH, Free Bend, Stretch Bend, Draw Bead (minor strain is zero), Olsen, Erichsen, Four-inch Dome, Marciniak (minor strain is positive) SINGLE MODE DRAW LDR, Swift Flat Bottom COMBINATION STRETCH PLUS DRAW Fukui, Swift Round Bottom

4.3.4 TEST DESCRIPTIONS 4.3.4.1 Hole Expansion


The hole expansion test measures the ability of a sheet edge to be elongated. No standard test procedures have been developed for this test. Thus, various combinations of hole diameters and punch diameters, as well as different punch configurations, are used by different laboratories. Many test units in operation today have a common diameter of four inches (102mm). The diameter of the hole depends on the sheet metal being tested; a typical hole diameter of two inches (51mm) allows a maximum hole elongation of 100 percent. One typical test procedure (Figure 4.3-1) starts with either an eight-inch (203mm) diameter circle or square blank placed in the test apparatus and a crimp or lock bead formed. The locking is required to insure that all deformation is confined to the enlargement of the hole and that metal does not move into the die opening from the binder area. The blank is removed from the test apparatus and a central hole is placed in the blank; the crimp bead is used by the punching device to accurately center the hole. The amount of hole expansion depends on the quality of the central hole. A milled hole will expand further than a punched hole which contains damaged edges. A poorly punched will expand less than a good quality hole. The hole expansion test, therefore, will allow different quality holes to be used as one of the test variables. The problem becomes one of reproducing the specific quality of a hole from one test specimen to another. The blank containing the hole is reinserted into the test apparatus and expanded by the punch. Different punch geometries can be used. While the hemispherical punch is the most common, conical and flat-bottom punches have been used in some laboratories. All the punches, however, generate an edge strain path which approximates a tensile test. During the circumferential elongation of the hole, the hole edge is free to contract in the radial direction. This creates the characteristic positive major strain and negative minor strain. Hole expansion is continued until a predefined end point is reached. This usually is the onset of edge checks or edge notches. Deformation beyond this stage will rapidly lead to cracking. Traditionally, punch load measurements are not made for determining the end point since the onset of checking may not cause a sufficient drop in the punch load to terminate the test. For this reason, end point determination usually is done visually. Directionality of sheet metal properties (anisotropy) usually causes the hole to deviate from a circular shape. Thus a series of diametrical measurements are necessary to determine an average hole expansion. The formula for the hole expansion is
% H.E. = D f - DI DI x 100

where Df is the final hole diameter and Di is the initial hole diameter.

A number of papers have been written about the hole expansion test (K-42, D-6, D-8, W-14, N-3). Originally used for studies of container flanging, the hole expansion test has grown in popularity for evaluating higher strength steels and the ability of various steels to withstand blanking damage. Some general observations can be derived form the hole expansion tests: The primary factor which influences the hole expansion values is the condition of the initial hole. This is illustrated in Figure 4.3-2 where the expansion values for different hole qualities and different strengths of steel (with and without inclusion shape control) are presented. Substantially greater variability in the hole enlargements were observed with the drilled holes (initiation and propagation of cracks) than with punched holes (propagation only). The percent hole expansion is: directly related to: sheet thickness work hardening exponent, n total elongation normal anisotropy, rm minimum r value and, inversely related to: yield strength tensile strength hardness A good correlation was obtained between the stretch bend tests and the hole enlargement tests (D-2).

4.3.4.2 Yoshida Buckling (Handkerchief)


The Yoshida Buckling test (YBT) is a test to evaluate the susceptibility of sheet metal to buckling (Y-5, Y8, H-12, J-4, Y-9, S-46, H-31, L-2, M-13, G-5, S-15). In this test, a square sheet metal sample is clamped on one set of diagonal corners and elongated in a tensile test frame (Figure 4.3-3). As the specimen is elongated, a series of wrinkles or buckles are generated parallel to the direction of elongation. The degree of width contraction and height of the wrinkles are measures of buckling.

4.3.4.3 Limiting Dome Height (LDH)


The Limiting Dome Height (LDH) test is a modified hemispherical dome test (A-12). Instead of a fully clamped blank, strips of varying widths are clamped on end by a lock bead and then deformed with a four-inch (102mm) diameter hemispherical punch (Figure 4.3-4). The different widths generate different minor strains at failure. The height at failure (height at maximum load) can be plotted as a function of strip width or minor strain. LDH curves for two different steels are schematically shown in Figure 4.3-5. Superimposed on the LDH curves are the strain paths commonly followed by three traditional laboratory tests. The total height of the dome depend on two factors. The first is the maximum amount of strain the metal can withstand before failure. Obtained from the Forming Limit Diagram (FLD), this strain represents a limiting (ceiling) value. The second factor is how uniformly the strain is distributed without exceeding the ceiling. Since the dome height is proportional to the area under the strain distance curve, a high uniform strain will generate the largest dome height. The distribution of strain is governed both by the stretchability of the steel and the strain distribution characteristics of the blank-punch interface. These are also the characteristics which lead to good stretchability in a production stamping. Therefore, it is argued that the height in the LDH test can be related to the stretchability of production stampings.

The shape of the LDH curve approximates that of the FLD. However, the ability of the substrate and the interface to distribute strain in the presence of biaxial stress state will affect not only the height but also the shape of the curve. The most common application of the LDH test is to measure and report only the minimum of the curve LDH0. In this case it is necessary only to measure the specimen width; measurement of the minor strain is unnecessary since the minimum occurs at plane strain. Test procedures for the LDH test have not been standardized. Various researchers (G-9, A-30, M-23), as well as the North American Deep Draw Research Group of the American Society for Metals, have undertaken numerous cooperative programs to converge on the best procedure. The description which follows is intended only as a guide. To determine the LDH0, seven-inch (178mm) long specimens of the test metal with various widths are sheared. Five different width strips are tested, each strip being 1/8 inch (3mm) wider than the previous strip. The widths of the five strips are chosen to bracket the width which will generate a plane-strain condition the minimum of the curve. For uncoated, low carbon, mild steel this width usually is approximately five inches (127mm). Each blank is processed according to prescribed test procedures. Here testing philosophies diverge. One procedure dictates maximum cleaning and degreasing to remove ill oil and other surface contaminants. A second procedure requires the specimen to be tested as-received, since mill oil, prelubes, etc. affect both the test result and the press shop performance. A third procedure attempts to standardize the interface in order to measure only the steel stretchability. A low viscosity wash oil appears to provide a more standard condition than the various cleaning techniques. The strip is centered in the test fixture and securely clamped. Absolute clamping is imperative. In this respect, the tooling must be highly standardized to provide interchangeable data. The strip is deformed by the hemispherical punch at a typical speed of ten inches per minute (254mm/min). The punch travel at maximum punch load is the Limiting Dome Height and is determined from autographic records or electronic maximum load detectors. These data allow the minimum of the LDH-width curve to be constructed and LDH0 determined. Different procedures to eliminate test variability include the use of standardized samples and seasoning of the punch after changing material type (m-23). The tool material can also be changed to duplicate soft tryout tools or hardened tools (B-16, M-24, B-17). The common tensile properties which contribute to good stretchability also contribute to high LDH values: work hardening exponent (n), strain rate hardening (m). total elongation, tensile/yield ratio, etc. Experiments indicate that the high mean anisotropy ratio (rm) important to cup drawing can be detrimental to the LDH value (H-21, N-12). the exact role of surface topography in controlling dome height needs further research. Some complex interactions of sheet surface material, sheet surface profile, lubricant, tool material, tool surface profile, interface pressure, and interface temperatures are present (B-16, B-17).

4.3.4.4 Free Bend


Bend tests are different from most other sheet metal forming tests in that a severe strain gradient is developed through the sheet thickness by the act of bending. A simple bend test (I-4) is made in one direction without reversing direction of bending (Figure 4.3-6). A more controlled variation of the bend test is to clamp a flat specimen against a die plate with a radius on one end and bend the strip to a specified angle by using a slowly applied force (Figure 4.3-7). Specifications may require the specimen used to be from either the rolling direction or the transverse direction of the sheet. The strip then is formed over a flat plate until its sides are parallel and a fixed distance apart relative to the strip thickness. This distance usually is given as a number times strip thickness, such as 4t for a four sheet thickness spacing. When required, no spacer may be specified and the strip is bent flat upon itself through a 180 degree bend

angle (Figure 4.3-8). This is called a 0t bend. In simple bend tests, the requirement is to obtain the specified angle or shape under load without regard to possible springback after the load is removed. The interpretation of results is a matter for the material specification. The metallurgical and mechanical factors important in all bending operations are the strength and ductility of the steel, the degree of inclusion shape control, and the condition of the edge of the test blank. Bendability is described by the radius to thickness ratio (in multiples of thicknesses) that the steel can be bent without developing cracks over a specified length. This, therefore, becomes a pass/fail system. Typical bend test data are shown in Figures 4.3-9 and 4.3-10. In Figure 4.3-9 poor correlation was obtained between bendability and percent total elongation (T-8). Better correlation was obtained between bendability and percent reduction of area. The effect of stringer inclusions for different yield strength steels is shown in Figure 4.3-10. A bendability definition which is superior to the pass-fail system is the length of the longest edge crack developed under constant bend conditions (K-55). This test combine outer fiber strain and the strain limits of the material. Using this criterion, the effect of sulfide shape control during bending has been assessed (Figure 4.3-11). The data in this graph shows the reduced bendability of an 80 ksi (550 MPa) steel compared to a 50 ksi (345 MPa) steel when neither steel has inclusion shape control. With inclusion shape control, both steels have equal bendability at a level significantly better than the 50 ksi (345 MPa) steel without inclusion shape control. Like many other simulative tests, bending tests are often used to evaluate the adhesion of various metallic and painted coatings.

4.3.4.5 Stretch Bend


Some sheet metal forming operations are pure bending operations. Other operations are pure stretching operations. The latter is true for thin sheets and generous radii, where the radius/thickness ratio (R/t) is greater than 15 and the bending component is considered to be negligible compared to the uniform strains through the thickness. Many production parts, however, are made from thicker steels or, more commonly, have tighter radii in character lines, sharp punches, embossments, etc. The Stretch Band test attempts to duplicate the combination of stretch plus bending. Two types of Stretch Bend tests have been suggested in the literature (D-11). The first is called the Angular Stretch-Bend Test (ASBT) shown in Figure 4.3-12. The angular punch has a single plane of symmetry. The rectangular blank used in the ASBT is eight inches (203mm) across the die opening and four inches (102mm) wide. A series of punches with different radii allow different ratios of R/t to be generated. More combinations can be obtained without increasing the number of punches by varying the thickness of the steel to be tested. The blank is firmly clamped and stretched. The depth of punch travel (configuration height H) is the measure of stretch-bendability. The test conditions are monitored to insure that fracture occurs over the punch. Fracture in the unsupported region of the strip would simulate tensile failure instead of the desired stretch-bend mode. The second Stretch Bend test, a variation of the angular punch test, is the hemispherical punch test (HSBT). Here the punch has a double curvature. In this case, blanks of varying width are used to provide varying amounts of drawing-in (such as the Limiting Dome Height test). The test results depend on the steels tested (Figure 4.3-13) and the test parameter R/t (Figure 4.3-14). The test results should be taken as a relative ranking of various steels rather than an absolute value which can be translated into design parameters for die construction. Another combination test has been proposed by Rasmussen (R-8) whereby the shearing properties of the sheet in the direction of the prior bending axis are used to indicate the severity of the bend and its resistance to subsequent deformation.

4.3.4.6 Draw Bead


Draw Bead tests are used to simulate the deformation a sheet of metal undergoes as it moves through one or more draw beads located in the binder (blankholder) area of the die system. Two very distinct type of Draw Bead tests are in use. One is used to test the adherence of coatings, while the other is used to measure the coefficient of friction of sheet metal passing through the beads. The coating adherence draw bead test is not a true sheet metal formability test. In these tests, the male bead and the female recess (Figure 4.3-15a) have identical radii. Any sheet metal inserted between the bead and the recess is heavily loaded at each shoulder due to the negative clearance which exists. The sheet metal is dragged (even scraped) through the beads. Visual examination of the surface, or a weight loss, is made after, typically, two passes through the beads. The second type of Draw Bead test is called the Draw Bead Simulator (DBS) because it attempts to simulate the deformation of the sheet metal as it passes through an ideal bead (N-14, N-15). Here a coefficient of friction is calculated for each combination of sheet, lubricant, and bead metal. The interchangeable beads (Figure 4.3-15b) can vary from hardened steel to a soft zinc-based alloy used for die tryout. A positive clearance (gap dimension greater than metal thickness) is used to avoid any pinching or scraping action. The normal force on the bead is not controlled but allowed to vary. Instead, the beads are interleafed a fixed amount and then locked. The pulling rate generally exceeds 50 inches per minute (1270mm/min) to duplicate more closely actual press speeds. The measured force to pull the specimen through a fixed bead has two components. One is the frictional force, which is used to calculate the coefficient of friction. The other is the force required to bend and unbend the sample as it passes around the beads. This force is obtained separately by pulling a second sample of the metal around frictionless roller beads. The force from the roller beads is subtracted from the force from the fixed beads to calculate a coefficient of friction. Studies to date (B-16, N-14, N-15, K-16, K-17, B-17, W-13) have shown a unique value of the coefficient of friction for each combination of steel surface morphology, interface lubricant, and bead metal. One test result can not be extrapolated to any other test combination each combination must be tested. The coefficient of friction generally increases as the test speed is decreased (K-17).

4.3.4.7 Olsen/Erichsen Dome


The Olsen Dome test is used as a stretchability evaluation test in North America. The Erichsen Dome test is a similar test used in Europe and Japan. The simple test procedures makes these two tests popular in the sheet metal forming industry. A good correlation has been found between these two tests (N-12). Therefore, comments will be restricted here to the Olsen Dome test. The operational procedures of the Olsen Dome test are few (A-26). Specimen preparation requires only sharing a small rectangular blank. The test apparatus consists of a hydraulic cylinder to force a small diameter hemispherical punch into a sheet clamped by an annular die set activated by another hydraulic cylinder (Figure 4.3-16). Measurements are made of punch load and punch travel. Punch travel is stopped when the end-point is detected. The correct end-point is the maximum in the punch load-punch travel curve. This can be determined from a graphical plot of punch load versus punch travel. Some of the more advanced Olsen Dome test machines are equipped with automatic, electronic, load maximum detectors. The common practice for most Olsen tests, however, is to observe the load indicator dial for the load maximum and then to manually stop the punch travel. The load maximum is sometimes difficult to detect accurately. Therefore, some operators observe the test specimen for a visual onset of the necking failure. Total punch travel is measured in thousandths of an inch; the Olsen value is the dome height in thousandths. Therefore, a 425 Olsen value is really 0.425 inches of punch travel to the point of test cutoff.

The degree of necking and the response time for machine shutdown vary from operator to operator and even for repetitive tests with the same operator. This difference, especially for inter-laboratory correlations, can be as large as 60 Olsen units (A-7). In theory the Olsen Dome test is a pure biaxial stretch test. As such the Olsen values should correlate with common stretchability parameters, especially the work hardening exponent (n). Available data, however, show extreme scatter (H-18, T-8, N-12, K-43). Little correlation is seen as a function of the n value (Figure 4.3-17). or the uniform elongation (Figure 4.3-18). Somewhat better correlation is obtained with percent transverse reduction in area (Figure 4.3-18). Correlation between the Olsen values and n were obtained only when oiled polyethylene sheets were used as a barrier lubricant indicating the extreme importance of frictional factors. Research studies have highlighted a number of weaknesses in the Olsen Dome test which can account for the lack of correlation between the Olsen value and the work hardening exponent, n (A-7, T-8, H-18, K-27):
a. Test parameters have been insufficiently defined. These parameters include absolute determination of end point, lubrication procedures, sufficient clamping, and speed control. For example, the Olsen Dome test is designed to be a pure biaxial stretch test. However, many test fixtures have insufficient clamping forces which allow flange metal to flow over the die radius and into the die cavity. The uncontrolled and predictable amount of metal flow randomly adds to the maximum height of the dome at failure and therefore can be observed by deviation of the edge of the test specimen from a straight line. A speed increase from 0.05 to 1.5 inches per minute (1.25 to 38mm/min) can produce an increase of 12 Olsen units for an automotive gage AKDQ steel sheet (K-27). In practice, the typical Olsen Dome test is performed at a very rapid speed until fracture is approached and then the speed until fracture is approached and then the speed is dramatically reduced in an attempt to read the dome height at fracture. Thus, the test speed is usually unspecified and nonconstant. Even the test parameters are specified, there are problems. For example, even between 0.060 and 0.062 inch (1.52 and 1.58mm) the punch dimension changes discontinuously from 1.0 to 1.25 inches (25.4 to 28.6mm) in diameter to accommodate the thicker material. b. The small diameter punch can introduce a bending component when the punch radius/sheet thickness ratio becomes less than 10-15. Thus, a bending component (and a change in stress/strain state) is introduced when the sheet metal exceeds 0.04 inches (1 mm) in thickness. c. A relatively small area of the sheet metal is evaluated in any one test. d. Poor inter-laboratory correlations have been observed (A-7). e. Corrections are required when comparing metals of different thicknesses (F-8). The Olsen value increase 2.5 units for each 0.001 inch (0.025mm) increase in gage (F-8).

However, the Olsen Dome test can be very useful in evaluating certain parameters of the steel. For example, the Olsen Dome test often is used to show the presence of orange peel at high levels of strain prior to fracture. In addition, the type of fracture (ring versus highly directional) provides some indication of steel cleanliness. The Olsen Dome test also is utilized for determining the adhesion of metallic and paint coatings to a substrate during stretch deformation. For this application, a Scotch-Tape peel test is usually included. However, even for these applications, a severe strain gradient form the pole to the rim of this small diameter dome complicates the test.

4.3.4.8 Large Diameter Hemispherical Stretch


An improved dome stretching test has been used over the last 20 years to evaluate pure stretch forming (K-15, H-24, H-14, H-18, H-19). The typical diameter of the hemispherical punch is 4 inches (102mm), although diameters from 3 to 12 inches (76 to 305mm) have been used. These dome tests are designed to eliminate the problems enumerated for the Olsen/Erichsen tests described above. They differ from the Olsen/Erichsen test because:
a. The four-inch (102mm) diameter punch evaluates 16 times the area of the Olsen test. The larger

diameter also reduces the bending component. b. The die ring contains a lock bead to insure pure stretching. c. The fixture is a subpress designed to fit into a instrumented tensile test machine or is designed as a complete stand-alone unit. Most of these units are computer controlled in terms of test speed and data acquisition. The punch travel at failure is taken at the maximum load point. d. Test procedures are not standardized. However, within a given laboratory, standard cleaning and lubrication procedures have been established.

These large diameter dome tests are used for a variety of purposes, including evaluating the influence of steel substrate stretchability, coatings, lubricants, test speed, temperature, tool material, and tool surface coatings on the dome height. In the extreme, the rigid hemispherical punch has been replaced by hydraulic fluid (of fluid pressure simulated by a thick plug of rubber) to create maximum strain under conditions of balanced biaxial stress at the pole of the dome. Examples of the data obtained are shown in Figures 4.3-19 (H-18). The lack of correlation is due to the fact that stretchability is a function of the work hardening exponent (n) plus the strain rate hardening exponent (m). The value of (n + m) correlates with the total elongation when different classes of metals are studied. In Figure 4.3-19 the following property combinations are observed: Zn: low n, high m Steel: medium n, low m Brass: high n, zero m Aluminum: medium n, negative m Thus, the dome heights of the metals shown in Figure 4.3-19 correlate well with total elongation in Figure 4.3-20. Within a given class of metal, steel for example, correlations between the work hardening exponent (n) and dome height can be observed (Figure 4.3-21). Note here that the old terminology of cup height is used. The current terminology is cup height for cylindrical deep drawn cups and dome height for biaxially stretched, clamped sheets. In Figure 4.3-21, the dome heights for tests run with oiled polyethylene sheets are greater than those run without lubrication (dry). However, the scatter for the polyethylene lubricated domes is greater and the slope of the curve is reduced. Thus, the better lubricant tends to mask the effects of other variables such as material properties. While this masking has useful implications in the press shop, it reduces sensitivity of the simulative tests. Therefore, simulative tests should be designed to maximize the effect of the variable under study. In a like manner, the n values for steel, and therefore the dome heights, decrease as the yield strength of steel increases (Figure 4.3-22). High normal anisotropy or strain ratio, rm, (important for good cup drawability) has been shown to be detrimental to pure dome stretching (Figure 4.3-23). Here the circle arc elongation (eca) is related to the uniform elongation and therefore the n value of the material. Therefore, for a given eca (or n value) the dome height is inversely proportional to rm values. No explanation for this effect has been proposed. It may be a direct effect of substrate stretchability or an indirect result of interface lubrication effects. Hecker (H-18) could not duplicate these results. Other dome tests have been conducted which evaluate test speed (Figure 4.3-24). While the Large Diameter Hemispherical Stretch test has many advantages over the Olsen/Erichsen Dome test, it still retains the strain gradient and varying strain states from the pole to the rim of the dome. These point to point variations can cause problems in analysis of results. Application of the Large Diameter Hemispherical Stretch test to actual production problems has begun. Brazier (B-18) used the four-inch (012mm) dome test to establish permissible pocket depths permissible pocket depths were established by the ratio of the dome height to dome diameter for various metals.

4.3.4.9 Marciniak Stretch

The Marciniak Stretch test (M-9, G-35) is a modified dome test (Figure 4.3-25). It was designed to overcome the severe strain gradients developed by the traditional dome tests using a hemispherical punch. If a flat sheet of metal is simply clamped and a flat-bottom punch is pushed into the sheet, a very limited amount of stretch is possible, usually around the punch radius, before the strain level reaches failure and tearing occurs over the radius. The low level of strain that does occur under the flat (central) portion of the blank is balanced biaxial because stretching is equal in all directions. To increase the level of straining in the flat bottom, metal must slide over the punch radius and transmit increasing force to the metal in the flat bottom. Strain can not be allowed to localize in the radius. This is extremely difficult to accomplish with a single sheet of metal. In theory, the punch radius could be replaced with a series of ball bearings to generate a frictionless radius. A more practical solution is to insert a carrier blank between the test blank and the punch (see Figure 4.3-25). A central hole punched in the carrier blank easily expands allowing metal to slide over the punch radius with relatively small applied force. The metal of the test blank on top of the carrier blank rides with the carrier blank. The metal of the test blank within the circumference of the carrier blank hole elongates in all directions. Several specimen sequences are possible. The simplest is to use a square carrier sheet and test sheet with a minimum dimension of one inch (25mm) greater than the dimension of the lock bead. A common size is eight inches (203mm) for a four-inch (102mm) diameter punch. The carrier blank can be made from any stock steel or other metal with excellent formability; the test bank must fail before the edge of the enlarged hole in the carrier blank checks or tears. Carrier blank thickness is not specified, but must be on the same order as the test blank or else fracture can occur in the carrier blank. The test blank and carrier blank are mated and placed in the test fixture to be crimped. This insures proper carrier hole alignment. The carrier blank is removed for punching of the central hole and then replaced. A two-inch (50mm) hole is common. The punch is driven into the pair of blanks (Figure 4.3-25) until the desired end point is reached. For a given test configuration, the amount of strain in the flat bottom is a function of punch depth. Therefore, blanks with identical amounts of strain can be generated as prestrained samples for other tests. Another end point is fracture, this end point is best determined by visual monitoring during deformation. The Marciniak Stretch test is a specialized test used in only a few laboratories. Coated metals (both metallic and painted) can be formed to increasing levels of strain. The balanced biaxial strain state generates maximum increase in surface area. Changes in coating adhesion, ductility, and surface topography are several parameters easily measured. Because no contact is made with the punch in the central test area, print-through problems are eliminated. The specimens are flat for application of different visual tests. All the specimens can be painted after forming to simulate painting of actual deformed panels. This test also indicates the internal cleanliness of the substrate based on the nature of the incipient fracture. A long, straight, rolling direction, line fracture with no necking elsewhere in the surface indicates a large inclusion. A surface with deep wormy necks in all directions over the entire surface indicates a very clean substrate. By evaluating samples from edge to edge of the sheet, a profile of internal cleanliness can be generated. This also can correlated with height at fracture (average strain) for the various samples.

4.3.4.10 Limited Draw Ratio (LDR)


The tests described previously were simulative stretch tests. The Limiting Draw Ratio test is a single mode draw test. This test evaluates the ability of a sheet metal to be drawn into a cylindrical cup (Figure 4.3-26). Test procedures have not been standardized (H-34, W-19, A-28, A-29). Typically, however, circular disks of increasing diameter are machined or punched. After deburring, they are inserted into the

draw fixture. A clamping force is applied which is sufficient to prevent wrinkling but which is not so large as to add an unnecessarily large component to the draw load. Cups are drawn with increasing blank diameter until the blank diameter is reached at which breakage is encountered. The Limiting Draw Ratio is defined as that maximum ratio of blank diameter to punch diameter for the onset of breakage. The absolute value of the LDR for any given lot of metal depends on die and punch geometry (Figures 4.3-27 and 4.3-28), test speed (Figure 4.3-29), temperature (Figure 4.3-30), lubrication, and holddown parameters. Once these are fixed, however, various metals can be ranked according to their relative capacity for deep drawing. The LDR increases with increasing sheet thickness (Figure 4.3-31). The LDR has been shown to depend linearly on the normal anisotropy of the steel (Figure 4.3-32). The mean Youngs modulus can also be used to predict the LDR (M-35). When evaluated over a wider range of materials, however, the LDR-rm relationship has been shown to be a linear relation on a log-log plot (Figure 4.3-33); this figure by Atkinson and Maclean (A-28) is popular in many paper textbooks. Hosford, however, argues that b.c.c. and h.c.p. metals should not be compared on the same line (H-37). When comparing only b.c.c. metals, the curve sharply decreases slope for increasing rm (Figure 4.3-34). He contends that very large values of rm are not beneficial in increasing the LDR. One Japanese author (Y10) claims that limiting depth is better defined by their X value rather than the rm values for a wide range of metals (Figure 4.3-35); the X value is a complex term incorporating work hardening anisotropy and changes in the yield locus. There is a one-blank method for evaluating deep drawability which greatly reduces blank preparation (A29, W-21). A blank of a given size is drawn into a cup of specified dimensions until the maximum load is exceeded. The blank is then clamped and the load to fracture the cup wall is measured. The greater the ratio of wall fracture load to draw load (Lu/Ld), the greater the deep drawability of the metal. This ratio can be empirically converted to a rm value of the metal (Figure 4.3-36).

4.3.4.11 Fukui Conical Cup


Very few stampings can be identified as being pure stretch or pure draw. Therefore, it can be argued that laboratory simulative tests which are intended to be pure stretch (such as the Olsen Dome test) or pure cup drawing tests (such as the Limiting Draw Ratio) should correlate poorly with actual press performance. A laboratory simulative test is needed which incorporates both modes of deformation. Fukui Conical Cup test was designed to overcome this problem (F-9). In this test, a circular disk of metal is blanked in a separate operation; its diameter depends on the sheet metal thickness. This disk is then placed into a conical die (Figure 4.3-37). No holddown is used. This is possible because the ratio of blank diameter to sheet thickness is such that buckling does not occur. The absence of holddown eliminates many of the test variables associated with draw type tests, such as holddown load, die radii, roughness of the holddown surfaces, lubrication under the holddown, etc. The deformation forces are generated by pushing a spherical ball into the center of the blank. Two deformation modes occur. The central portion of the blank is stretched over the ball. This loading also causes the blank to be pulled down the conical cavity. The circumference of the blank must decrease, generating the compressive stress component found in cup drawing. This test sums all the various components to produce a final cup height at failure. The higher the maximum height, the more formable the steel. The traditional Fukui Conical Cup Values (CCV) the ratio of final cup diameter to blank diameter strongly depend on the product of nm times rm (Figure 4.3-38). The change in Fukui CCV values for AK steel as a function of test speed is shown in Figure 4.3-39. The end point of this test was modified by Goodwin and is called the Formability Index (G-24). The punch travel at maximum load is the Formability Index; in this respect, the modified Fukui test is similar to the Olsen Dome test. Goodwin showed that the Formability Index is related to the product of the minimum r value and the uniform

elongation of the sample tested (G-24). The minimum r value for cold rolled steel usually occurs in the 45 degree direction to the rolling direction.

4.3.4.12 Swift Round Bottom Cup


The original Swift Flat Bottom test (C-14) was a pure cup drawing test. To make the test sensitive to both stretch and draw components, the flat bottom of the punch was replaced by a hemispherical head, typically four inches (102mm) in diameter (Figure 4.3-40). The test is conducted similarly to the Limiting Draw Ratio test. Blanks of increasing diameter are formed with the punch until breakage occurs.

4.3.5 SUMMARY
A wide variety of simulative tests are available in the literature. Some are well defined with all variables and test parameters specified. Others are essentially undefined with equipment and procedures varying from one investigator to another. Each test represents an attempt to duplicate some portion of a complex stamping or a specific forming mode. Many are successful; others are not. As the critical forming mode in any given stamping changes, so the required test changes. The current trend is to increase the number of simulative tests performed in order to broaden the characterization of a specific metal or lubricant. In addition, the simulative tests provide excellent data for verification of mathematical simulation models. Figure 4.3-1 Schematic for the Hole Expansion test. Specimen after an increase of deformation is shown as dotted lines.

Figure 4.3-2 Hole expansion results are strongly influenced by edge condition. Steels indicated by l and l' were inclusion shape controlled steels. The 30, 50, 60, and 80 numbers refer to minimum

yield strengths in ksi (H-29).

Figure 4.3-3 Yoshida Buckling test specimen (H-12).

Figure 4.3-4 Schematic for the Limiting Dome Height test. Specimen after an increment of

deformation is shown as dotted lines.

Figure 4.3-5 Schematic showing Limiting Dome Height (LDH) for two different metals. The minimum is labeled LDH0 (A-12).

Figure 4.3-6 Simple bend test for sheet metals (A-12).

Figure 4.3-7 Figure 4.3-7 Modified simple bend test. (a) is start of bend. (b) is finish flat using specified spacer designated as 1, 2, 3-t (times sheet thickness) (A-12).

Figure 4.3-8 Bend flat upon itself is the most severe test condition - referred to as 0-t (zero thickness) (A-12).

Figure 4.3-9 Bendability as a function of total elongation and transverse reduction in area for

various strength steels (T-8).

Figure 4.3-10 Stringer Inclusions reduce bendability (T-8).

Figure 4.3-11 Using length of edge cracking as a measure of formability (K-55).

Figure 4.3-12 Schematic of the Stretch Bend test. Specimen after an increment of deformation is shown by the dotted lines.

Figure 4.3-13 Stretch-bent height as a function of yield strength (F-2).

Figure 4.3-14 Transverse stretch-band tests with and without rare earth treatment (F-6).

Figure 4.3-15 Schematic showing two types of Draw Bead tests. The coating adhesion test (left) heavily loads the strip with a negative clearance. The coefficient of friction test (right) measures pulling load for friction plus deformation (fixed bead) and pulling load for deformation only (roller

bead).

Figure 4.3-16 Schematic of Olsen Dome test. Specimen after an increment of deformation is shown as dotted lines.

Figure 4.3-17 Olsen Dome height does not correlate well with the work hardening exponent, n (H18).

Figure 4.3-18 Olsen Dome height correlates with transverse percent reduction in area (T-8).

Figure 4.3-19 Poor correlation of stretch dome height with the work hardening exponent, n, exists for different metals.

Figure 4.3-20 The stretch dome height correlates with total elongation (H-18).

Figure 4.3-21 Dome heights for large diameter domes formed with and without lubrication (H-18).

Figure 4.3-22 Maximum dome height of a round bottom dome is dependent on the work hardening

exponent, n (F-2).

Figure 4.3-23 Dome height as a function of circle arc elongation (N-12, H-21).

Figure 4.3-24 Dome height at failure as a function of punch speed (G-10).

Figure 4.3-25 Schematic of Marciniak Cup test. Specimen after an increment of deformation is shown as dotted lines.

Figure 4.3-26 Schematic of Limiting Draw Ratio (LDR) test. Specimen after an increment of

deformation is shown as dotted lines.

Figure 4.3-27 Limiting Draw Ratio (LDR) decreases as sheet thickness decreases (F-8).

Figure 4.3-28 Maximum Draw Ratio (LDR) as a function of both the die and punch profile radii (H-

34).

Figure 4.3-29 Critical blank diameter for flat bottom punches increases with punch speed (F-7).

Figure 4.3-30 Limiting Drawing Ratio (LDR) decreases as the punch temperature increases (G-32).

Figure 4.3-31 Recommended Limiting Drawing Ratio (LDR) as a function of sheet thickness (J-13).

Figure 4.3-32 Limiting Drawing Ratio (LDR) as a function of both rm and Em (M-35).

Figure 4.3-33 Linear relationship found between Limiting Draw Ratio (LDR) and rm for a wide variety of metals (A-28).

Figure 4.3-34 Limiting Draw Ratios (LDR) as a non-linear function of rm for a variety of metals as published by Hosford (H-37).

Figure 4.3-35 Relationship between limiting depth and X and r values (Y-10)

Figure 4.3-36 Approximation of the rm value from a single blank cup draw test (A-29).

Figure 4.3-37 Schematic of the Fukui Conical Cup test. Specimen after an increment of

deformation is shown by the dotted line.

Figure 4.3-38 The diameter ratio of the Fukui conical cup test correlates well with the product of nm and rm (F-8).

Figure 4.3-39 The diameter ratio of the Fukui conical cup test for various punch speeds and test

steels (F-7).

Figure 4.3-40 Schematic of Swift Round Bottom Cup Test.

4.4 Coated Steels


4.4.1 INTRODUCTION
The use of coated steels for automotive applications is rapidly increasing. Starting simply many years ago with galvanized rocker panels, the current vehicle designs are candidates for a wide spectrum of coated products ranging from prepainted cold-rolled steels on one extreme to multi-layers of different electrodeposited metals to the other.

The introduction of coated products into the press shop was not been without problems. Little press shop experience with these new products, if any, has been available to the artisan. Therefore, the artisan has been forced to implement the traditional press shop techniques of trial and error. Sheets of a new coated product have been inserted into available tooling and their press performance compared with that of the existing bare steel sheets. The results have been mixed at best and confusing to most. Some of the products have been so slippery that restraint in the binder has been impossible and the resulting stampings were a series of buckles, waves, and loose metal. Other products have welded to the binder and punch and would not flow into the die under any conditions; stampings were removed from the tooling in multiple pieces. When compared with uncoated steel sheets, many coated steels have different characteristics which affect formability in many unpredictable ways. Attempts at characterizing the formability of coated products by traditional mechanical properties have explained some of the differences in formability. Depending on the process used to make the product, different values of the substrate properties have been obtained. These in turn have influenced the press performance of the coated steels, just as the formability of bare steels would be affected by the same changes in substrate properties. This has explained some of the variations in press performance of galvanized steels. However, press shop experience has also shown that the press performance of some coated products has varied widely even for steels with identical substrate properties. This has confirmed the long held suspicion that the coatings also have had a major influence on the press formability of the product. Previous discussions of formability in this document have emphasized the importance of the four components of the Forming System: material, lubricant, die design, and press. Therefore, this section on coated steels also will review formability with respect to the Forming system. Coated products can be directly related to two of the four components material and lubricant. One approach to analyzing coated products is to separate the basic formability of the substrate from the effect of the coating (K-13, K-22). In this model, the formability of the substrate, as defined by its mechanical properties, determines the primary formability of the coated product. The mechanical properties of the substrate steel like uncoated steels determine the ability of the coated steel to withstand strain in the various modes of forming (A-13). The coating, in turn, affects the amount of the metal flow over the tool and die surfaces. In this manner, the effect of the coating parallels that of a lubricant. In this section, the coating on the steel, the lubricant in the workpiece-tool interface, and the surface of the tool are considered as a single system interacting together to control the flow of metal over the various tool surfaces. The coefficients of friction are the measured output of the system. This separation of the substrate formability from the lubricity effects of the coating presupposes no interaction between the coating and the substrate formability. Some research (S-36, S-37) has suggested that some coatings can reduce the formability of the substrate. For example, the zinc-iron layer formed at the surface of the steel substrate of a hot-dipped galvanized product is said to cause local tensile stresses which lead to premature failure in regions of negative curvature which contact the punch. This model is used to explain some observations that the limiting dome height in punch stretching decreases with increasing intermetallic layer thickness (S-37). A number of other studies, however, have been conducted comparing the formability of coated and uncoated steels which have led to the opposite conclusion (M-25, M-24). To insure identical substrate formability, the uncoated steel conditions are obtained by stripping the coating off the coated steel; the tool-surface effects are eliminated by isolating the sheet surfaces from the tool surface with an oiled polyethylene sheet. Meuleman, Denner, and Cheng (M-25) have shown that for both plane strain and stretch deformation modes, the zinc coatings had a negligible effect on the formability. In terms of drawability, only hot-dipped zinc-iron alloy coatings exhibited decreased drawability parameters relative to

uncoated steels i.e., reduced rm values. However, these reduced rm values can be explained as an artifact of the tensile test method rm value determination. The corresponding reduction in limiting draw ratios has not been observed (M-26). Additional research is required in this area. Finally, the position can be taken that any failure of the coating itself during the forming process even without substrate breakage should be considered a formability problem because the coating failures are affected by the mode and amount of the deformation. This category includes lack of adhesion, decohesion, scoring, galling, and other problems. The discussion following, therefore, is divided into three topics: - Substrate formability - Interface friction - Coating failures The primary emphasis of the discussion will be various types of galvanized coatings on steel substrates. However, the discussion is general and is equally applicable to other metallic coatings and paints, as well as substrates other than steel.

4.4.2 SUBSTRATE FORMABILITY


The formability of the substrate can be measured by the traditional formability parameters, including the work hardening exponent (n), strain rate hardening exponent (m), plastic anisotropy ratio (r), uniform elongation, total elongation, and other mechanical properties. Specific combinations of steel composition and processing conditions allow a wide range of different formability parameters to be developed in the steel substrate in order to meet the formability requirements of different stampings. A full discussion of these parameters is provided in Section 4.1. Historically, the galvanized steels used initially for the automotive industry were produced by the hotdipped galvanizing process. Two different types of hot-dipped galvanizing processes have evolved (Figure 4.4-1). One hot-dipped galvanizing process uses cold-rolled steel which has been pre-boxed annealed to obtain a soft, ductile structure that exhibits good formability. This steel then is heated to a temperature range of 850 to 900 degrees F (455 to 483 degrees C) to ensure that the steel is at the same temperature as the molten zinc bath. This low temperature process also is known as the Cook-Norteman process. The other process begins with the fully cold worked steel obtained from the tandem mill. This steel is heated to a temperature range of 1250 to 1600 degrees F (695 to 970 C) to achieve in-line annealing to replace the box annealing cycle. This high temperature process is known as the Sendzimir process. Because the in-line annealing is short in duration compared to box annealing, a slightly less formable steel results (A-13). The rapid heat up does not allow for the recrystallization and growth of crystallographic textures which lead to high normal anisotropy, rm (B-10). The rapid heating and the rapid cooling of the strip from the annealing furnace temperature to the temperature of the galvanizing pot causes excess carbon to remain in solution in the steel and a smaller grain size; these reduce the stretchability of the steel through lower values of the work hardening exponent, n. Correspondingly, the yield and tensile strength values are elevated and the total elongation is reduced. For these reasons hot-dipped galvanized steels as a class have been considered to have inferior formability. The electrogalvanizing process eliminates the heating and cooling of the steel required by the hot-dipped galvanizing process. For this steel, the electrogalvanizing step is added at the end of the normal coldrolled steel processing cycle. The formability of the substrate of the electrogalvanized steel therefore should be identical to the formability of its cold-rolled steel counterpart.

To improve the formability of the hot-dipped galvanized steel substrate, the steel producers have developed a number of processing options (Figure 4.4.2). The first option is a post batch anneal. While the rm value is unchanged, a slight increase in the n value and stretchability is achieved. The major change is in the lower yield strength. The second option is both a pre and post anneal in a batch anneal furnace. The properties now are similar to the bare cold-rolled steel. Superior stretchability can be achieved through special chemistry and processing. The most common here are ultra-low carbon steels (with or without additions) processed by vacuum degassing. The result of these chemistry/processing options is that the galvanized steel product received by a stamping plant may have different formability characteristics depending on the specific route a particular supplier chooses to meet the formability requirements of each stamping. However, as illustrated schematically in Figure 4.4-3, the steel producers have sufficient options available to them with which to produce one or more types of galvanized steels with the formability equal to or even exceeding that available with bare, cold-rolled steel (A-13). Interestingly, one study (G-38) has shown that the zinc layer improved the formability of steel in the stretching area by raising the FLD. The argument is made that the zinc increase the total thickness of the sheet ad therefore also increases the level of the FLD.

4.4.3 INTERFACE FRICTION


An obvious difference between bare steel and coated steel and coated steel is the frictional effect the coating has on the interface. The coefficient of friction is the resultant of a specific combination of workpiece (coated sheet metal), interface lubricant, and tool surface. Coating the sheet metal can be considered as adding another component (another layer) to the interface lubricant system. Recent research has shown that no single test can evaluate coated steel/lubricant combinations (M-25, K13, B-16, B-22, A-13, K-17, M-24, K-16, R-12). Instead, a variety of tests are required which will simulate various forming modes. This is illustrated in Figure 4.4-4. The specific forming mode is more important to coated steels than uncoated steels. In addition to the specific response of the steel substrate to the forming mode, the coating will respond differently to each strain state. One extreme comparison would be cup drawing versus biaxial stretching. In cup drawing the surface area of the flange is decreasing. Here the coating does not have to flow to provide coverage during the generation of new surface area of the substrate. However, the coating is subjected to compressive stresses which could cause a high strength coating to buckle (as opposed to upsetting upon itself) and separate from the substrate. The opposite occurs in biaxial stretching. Here the deformation is tensile which prevents coating buckling. However, if the ductility of the coating is less than the ductility of the substrate, then the coating will crack and create voids in the coating. Another deformation mode occurs in the draw bead area of the stamping. This deformation is plane strain bend-and-straighten. However, the coating is on the outer surface of the bend and therefore is subjected to the maximum strain. Even worse, each surface is subject to alternating tension and compression cycles which are demanding both on coating ductility and adhesion. A single coefficient of friction for each coated steel/lubrication combination independent of forming mode would be ideal for the press shop. However, recent research has shown the opposite (M-25, K-17, M24). The coefficient of friction for each combination of coated steel and lubricant depends on the specific forming mode to which the combination is subjected. Even worse, the rank order of a lubricant changes with variations in steel coating and forming mode. Likewise, for a given steel coating, the rank order of different lubricants changes for different forming modes. Finally, the coefficients of friction are further modified by forming speed, interface pressure, interface temperature, and other forming process variables.

Laboratory prediction of forming performance of coated steel products currently can be accomplished only for a specific set of test parameters. Thus, one laboratory test will be applicable only to one small segment of any complex production stamping. Therefore, current laboratory evaluation of coated steels are even more restricted in scope and applicability than laboratory evaluation of uncoated steels. The avenues for solving this complex problem appear possible, but need further research to develop press shop feasibility. First, identify test procedures which will provide a significant coefficient of friction for each forming mode which will encompass a large population of steels. For example, the coefficient of friction from a punch radius test conducted at one inch (25mm) per minute may correlate with the severity of most large, rectangular boxes, while a coefficient of friction from a high speed dome test will correlate with pure stretch forming such as a door handle pocket. Thus, one test could be sufficient to predict press shop behavior, to a first approximation. Second, perfect mathematical modeling and other research such that all forming conditions can be theoretically derived or empirically calculated from one or two key frictional tests. Third, make stampings insensitive to various types of steel when using a restricted number of mill applied lubricants and then prohibit use of in-plant lubricants. This restricts the number of possible combinations, and therefore restricts variations from a single source supplier. The problem is compounded when the tool material is changed. One common example is the soft, zincbased alloys used for die tryout and prototyping. The effect of the zinc coating depends on the deformation mode. For example, a hot-dipped, zinc-iron alloy, coated steel, commonly known as galvannealed steel, shows consistently diminished formability regardless of the deformation mode when tested with soft, zinc-based tools (M-24). In contrast, other hot-dipped free zinc and electrogalvanized steels show decreased formability with soft tooling only when substantial sliding of the blank occurs over the tooling, such as with soft draw beads; these same materials show improved performance in planestrain stretching with a soft, zinc-based punch. Thus, the galvannealed and the free zinc coated steels have performance changes in the opposite directions for punch stretching, but in the same direction for metal movement in the binder area. Thus even relative performance ranking between these various galvanized coatings is lost when changing from soft, zinc-alloy based tooling to hard, steel tooling. The reduced formability of the hot-dipped galvannealed steel is sufficiently severe that both Meuleman and Brazier (M-22) suggested that soft tool tryouts with galvannealed steel were not representative of how the galvannealed steel would perform in hard tools. These two authors independently concluded that soft tool tryout with bare steel (with comparable substrate properties) would best duplicate the forming conditions of the galvannealed steel in hard tooling. Most of the literature addresses problems with galvanized steels. However, identical analyses can be performed on prepainted steels (W-11). In summary, the coated steel-lubricant-tool interaction is so complex that no specific guidelines can be provided here. This conclusion has a significant impact on the mathematical modeling of coated steels, since many mathematical models are sensitive to the coefficients of friction for accurate calculations.

4.4.4 COATING FAILURES


All of the previous discussion has assumed that the coating has maintained its integrity, has remained bonded to the steel substrate, and has not welded itself to the tool. Some coatings do not respond in such a predictable, steady-state manner. These coating failures are yet another type of stamping rejection which depend both on the forming mode and the level of deformation. This concept is elucidated in a paper by Sudoo, Hayashi, and Nishihara (S-43) where a Flaking Limit Diagram is used to define deformation behavior of surface films (Figure 4.4-5). The axes for this diagram

are the same as the Japanese Forming Limit Diagrams, which differ from other Forming Limit Diagrams in that the axes are rotated first 180 degrees around the y axis (mirror image) and then 90 degrees clockwise. This paper concluded:
a. The flaking limit curve is remarkable different for different coated steels. The deformation mode favorable to one coated steel is sometimes unfavorable to another coated steel.

b.

Galvanized steel formed under biaxial stretching is sometimes susceptible to flaking. This is accelerated by a shear stress at the boundary of the surface film. c. In contrast with galvanized steel, galvannealed steel flakes easily in shrink flanging. Even though the strong bonding of the galvannealed film is useful in preventing flaking in biaxial stretching, the brittleness of this film causes easy breaking and flaking under compressive stresses which causes the coating to drop off as powder.
d. Zinc-rich primer coated steel is susceptible to flaking in biaxial stretching and shrink flanging. Flaking of this coating is considered to consist of both dropping off of zinc particles (cohesive failure) and exfoliation of the film itself (adhesive failure).

Shiokawa et al (S-20) performed a cylindrical cup test or a hat channel drawing test (including pre and post specimen weighing) in order to standardize an evaluation procedure for powdering. Their tests showed wide differences between different coatings, or even variations of similar coatings, in terms of weight losses due to powdering. Other papers confirm the complex nature of coating failures (H-36, N-11, E-4). The nature of coating deformation was examined in more detail by Makimattila and Ranta-Eskola (M-2). They found that the biaxial stretching of galvanized steels can be subdivided into two stages. In the first stage cracks nucleate and grow in grains that have brittle crystallographic directions oriented unfavorably with respect to slip directions; this is referred to as partially plastic deformation. After a gradual transition, deformation is characterized by more brittle behavior. Primary cracks widen and secondary cracks nucleate as the base steel is strained. Schedin, Karlson, and Melander (S-10) suggest a plasticity index (k) for coatings, where k = 0 for a coating which does not deform plastically and k = 1 for a coating which deforms as the base metal without cracking. A value of 0.7 was obtained for unixial tension, plane stain, and equibiaxial stretching for a commercially produced hot-dipped galvanized product. A two-stage coating evaluation test has been suggested (M-2). The first deformation mode is a biaxial stretching test. If the coatings pass the biaxial stretching test, they are then subjected to the more severe bending test. Other tests proposed to study the adhesion and cohesion of coated steels are they cylindrical cup test and the beaded hat channel test (S-20, E-4). Based on an evaluation of different combinations of factors-flaking, powdering, frictional resistance, and instability in subsequent stampings (die buildup and panel damage) various authors propose different types of galvanized steel as being the best for formability. Another type of coating failure is galling. Apparently galling begins when interface pressure exceeds some limit (M-2). However, once galling is initiated, the friction coefficient tends to decrease during sliding as built up particles on the tool surface become coarser and cause fewer contact points. A thicker galvanized coating tends towards higher galling limit pressure, reducing the risk of galling. Likewise, large draw beads and die radii are beneficial since changes in the surface topography due to bending deformation are small.

4.4.5 SUMMARY

One theme is constant in all the papers reviewed formability has many definitions, many modes, and many different types of failure. The coated steels have an added variable (the coating) which adds yet another dimension to the already complex matrix of interactions. This complex interaction makes the coated steels less predictable from stamping to stamping and from laboratory to press shop. In addition, when the stamping has a zero safety factor, the coated steels are even more susceptible than uncoated steels to the prevailing forming parameters. The only practical solution today is to model the galvanized steel in terms of the Forming System. The formability of coated steel is primarily dependent on the properties of the substrate and must be specified in the same manner as uncoated steel. The coating and lubricant combination must then be determined so as to provide metal flow patterns consistent with formability requirements of the stamping. This selection currently is guided by trial and error. Finally, the steel producer must user must implement steps to protect the more vulnerable coatings. Figure 4.4-1 Two process cycles for producing hot-dipped galvanized steel compared to traditional method for producing uncoated, cold-rolled steel.

Figure 4.4-2 A variety of special processing options can be employed to improve the formability of

hot-dipped galvanized steel (A-13).

Figure 4.4-3 The relative formability of the various galvanized steels encompasses the span of

formability of uncoated cold-rolled steels (A-13).

Figure 4.4-4 Different forming modes can be simulated by different simulative tests.

Figure 4.4-5 Flaking Limit Strain Diagrams for various galvanized steels (S-43).

4.5 HIGHER STRENGTH STEELS


4.5.1 INTRODUCTION
Formability of higher strength steels, especially the HSLA variety, is simply an extension of the formability of low strength steels. The same analyses are applicable, the same rules apply, and the same predictions can be made. The primary difference is that the specific values of the formability parameters are generally lower. Formability of higher strength steels depends on the material parameters reviewed in Sections 4.1 and 4.2. These include n, m, r, FLD, uniform elongation, and total elongation. The primary question to be asked for any higher strength steel is what the values of the formability parameters are when the minimum required yield (or tensile) strength has been achieved. These values will determine the relative formability of the steel under investigation compared to low strength steel, as well as compared to other higher strength steels which also meet the same strength requirements. In this respect, the

composition/processing combinations used to obtain the required strength are important only as they influence the formability parameters. For example, a 60 ksi (415 MPa) yield strength steel can be obtained through grain size control, alloying elements, or by cold work. Evaluating the formability parameters of the final product will indicate that cold work is not the best method to obtain the necessary yield strength because the formability parameters will be substantially lower, and therefore not competitive, with respect to steels strengthened by other techniques. If the steel is simply to be used in a shallow box with little forming requirement, then cost plays the dominant role. However, in automotive stampings, the maximum formability usually is required. The decision then is made on the level of the formability parameters and the cost necessary to obtain them.

4.5.2 FORMABILITY PARAMETERS 4.5.2.1 Work Hardening Exponent


The relationship between yield strength and the work hardening exponent, n, is shown in figure 4.5-1. For yield strength less the 45 ksi (315 MPa), little n value change is noted with the possibility of a large scatter band. Above 45 ksi (315 MPa) yield strength, the n value decreases linearly. This curve is useful for estimating the n value for any steel which has competitive formability. Steels which have strengths created by cold work have n values which lie below the curve and are not competitive. A German review of HSCR steels (K-2) contains a similar yield strength curve (Figure 4.5-2). The data here are below the middle line in Figure 4.5-1, indicating slightly lower n values for equal strengths. Another German paper (W-25) reviews four other steels. The dual-phase steel falls in the right extension of the curve in Figure 4.5-2 and the rephosphorized steel falls in the existing curve. The mild steel was given two skin passes: 1.5 and 3.0 percent. The 1.5 percent skin passed sample had an n value of 0.18 versus a range of 0.22 to 0.24 from the curve in Figure 4.5-2 for equivalent yield strength. The 3.0 percent skin passed sample had an n value of 0.16 versus a range of 0.20 to 0.22. Thus, even small amounts of skin passing, beyond elimination of yield point elongation, can generate large reductions in n value. A reduction of 0.05 in n value has a significant reduction in stretchability of the sheet metal. The n value versus yield strength curve is an excellent method of evaluating the expected stretchability of any given steel. In turn, the uniform elongation of steel sheet is directly related to its n value by the equation n = In (1+ uniform elongation) for steels which follow parabolic hardening.

4.5.2.2 Strain Rate Hardening


The post-uniform elongation can be related to the strain rate hardening exponent, m. The m value, in turn, is related to the strength of the steel; the m value decreases as the strength of the steel increases; this is shown in Figure 4.1-15 form reference S-9.

4.5.2.3 Total Elongation


Two major components of the total elongation are the uniform elongation (related to the n value) and the post-uniform elongation (related to the m value). Since both components decrease with increasing strength, the total elongation decreases with increasing strength (Figures 4.5-3 and 4.5-4). Figure 4.5-3 again shows the detrimental effect of achieving strength by cold work.

4.5.2.4 Plastic Anisotropy Ration


For most higher strength steels the plastic anisotropy ration, rm, is near unity. Typical data are shown in Figure 4.5-5. The rm values of micro-alloyed cold-rolled sheets ranges from 0.8 to 1.3. The

rephosphorized steels (P275), on the other hand, have rm values ranging from 1.3 to 1.7; these steels have good deep-drawability and could effectively compete in formation of oil pans, inner door panels, etc.

4.5.2.5 Forming Limit Diagrams


The effect of work hardening (inverse effect of strength) on the Forming Limit Diagrams is shown in the FLD0 nomograph shown in Figure 4.5-6. No effect is noted for n values greater than 0.21. This probably is related to the n value yield strength effect noted previously in Figure 4.5-1, but no studies have been conducted in this area. For n values less than 0.21, the reduction in the FLD0 with n value basically is a linear effect. The effect of sheet thickness in Figure 4.5-6 can be interpreted in two ways. First a 50 percent reduction in n reduces the FLD0 by 50 percent. This same reduction holds for all sheet thicknesses. Therefore, it could be argued that sheet thickness does not affect the strength FLD0 relationship. However, in absolute terms, 50 percent of 60 strain percent is much greater than 50 percent of 30 strain percent. Therefore, in terms of absolute strain percent reductions, the thicker steels suffer a greater loss in FLD0 as the strength increases. The dual-phase steels have provided interesting studies. Thompson and Hobbs (T-10) showed that the FLD0s for dual-phase steels are no different from other steels when compared in terms of uniform elongation. Keeler (K-33) showed that the FLD of a dual-phase was similar to that of a 100 ksi (695 MPa) yield strength steel. Thus, at onset of localized necking depicted by the FLD, the n value at the necking strain (often called the terminal n) is the important n value. Thus, at necking, the dual-phase steel and the 100 ksi (605 MPa) yield strength are similar in resistance to localized necking; previous strain characteristics are not important.

4.5.3 SPRINGBACK
The standard rule of thumb is that springback increases with increasing yield strength. This rule generally is experimentally verified with a simple bend test. Most automotive panels are not so simply deformed (Section 8.3). For example, studies of an outer side sill showed springback increasing with increasing strength. However, the increase was completely overshadowed by the reduction in springback due to restriking (Figure 4.5-7). The springback characteristics of dual-phase steel have generated research interest, since this steel is both a low strength and high strength steel during its forming history. Nakagawa and Abe (N-2) report small springback for small bending curvature because of the low yield strength; large springback is reported for large curvature because of the high work hardenability.

4.5.4 FORMING EXPERIENCE


The literature contains few, well documented case histories on the relative formability of higher strength steels. The formability of higher strength steels is a complex interaction of all variables. One may be tempted to base formability analysis only on the reduction in the FLD. However, peak strain levels also need to be considered; for example, if the current safety factor is 20 strain percent, a higher strength steel which reduces the FLD by 10 strain percent will not maintain the desired safety factor of 10 strain percent. The reason is that the FLD is both lowered with increased strength and also the peak strain is increased with increased strength (Figure 4.508). Thus, both effects must be considered. This is why simple knowledge of the change in the FLD with strength is insufficient to predict how a higher strength steel will perform in a stamping currently made with low strength steel. Too many unknown parameters enter into the creation of the strain distribution to predict accurately the peak strain. In one study (K-29), higher strength steels were placed in tooling designed for lower strength steel. The higher strength steel resisted deformation under the punch and transmitted a higher force to the material under the binder. This higher force overcame binder restraint forces and permitted more metal flow from the binder. This reduced the level of stretch required under the punch necessary to create the stamping depth. The higher strength steel could withstand less stretch under the punch, but in effect, the higher strength steel compensated for this stretch reduction by pulling relatively more metal from the binder area.

The increased strength of the steel sheet reduces the amount of stretch which can be induced in the center of automotive body panels. For this reason Asai et al (A-23) recommend that higher strength autobody panels for formed with stretch draw forming instead of conventional double-acting draw dies. This can increase the strain level by 50 to 100 percent of the normal level. The dent resistance for the stretch draw panels will be raised about 10 percent. Of course, the strain window for the higher strength steel is smaller and the stretch forming operation must be carefully controlled not to exceed allowable stretch limits. Similar process changes were recommended by Wollrab and Streidl (W-24); they encouraged increased blankholder forces or larger blanks to prevent wrinkling and surface deflections with higher strength steels. The higher pressures could result in increased tool wear, however. Similar results were found by Sato et al (S-7) who documented their work in Figure 4.5.9. However, as the strength of the steel increases, the ability to induce center panel straining by increased blank holder force is diminished. Press shop experience forming an intrusion beam from a 140 ksi (1000 MPa) tensile strength steel showed that while the maximum forming height is normally low for this steel, two drawing stages gave almost the same results as with mild steel formed under identical conditions (M-29). In this case the first stage forming was done with a large punch radius and the second with a small radius. A formability study done on higher strength, cold-rolled, sheet steels with a 58 ksi (400 MPa) tensile strength concluded that is was impossible to improve all properties to the level of mild steel sheets (S-18). Therefore, the applicability of specific higher strength steels should be considered with respect to the deformation mode of the intended application. The characteristics of the six steels studied are portrayed in a Formability Balance as shown in Figure 4.5-10.

4.5.5 SUMMARY
Generalized statements concerning the formability of higher strength steels sometimes are misleading or even incorrect. In terms of stretching, the work hardening exponent of the steel decreases with an increase in yield strength of the steel. This decrease is not a discontinuous loss of stretchability but is a gradual decrease well defined in the literature. On the other hand, the normal plastic anisotropy ratio is a function of steel processing and is independent of the yield strength per se. Sometimes the change in formability limits caused by different material properties is offset by a change in deformation over the tooling. Flow patterns usually will change because of the different material properties. The net deformation change, therefore, for higher strength steel may be more favorable than for lower strength steel. Springback in a pure bending configuration increases with increasing yield strength of the steel. However, deformation sequences are possible which will eliminate all the springback. In terms of formability, higher strength steels should be categorized as having different forming characteristics with no connotations attached as to whether these characteristics are good or bad. On this basis, the tooling can be designed and turned to accommodate these different characteristics to produce satisfactory stamping. Figure 4.5-1 A linear relationship exists between yield strength and n value for steels with a yield

strength greater than 45 ksi (315 MPa) (K-29).

Figure 4.5-2 Relationship between n value and yield strength (K-2).

Figure 4.5-3 The total elongation for a given yield strength depends upon the strengthening

mechanism (K-55).

Figure 4.5-4 Relationship between total elongation after fracture and yield strength of cold rolled steels for a series of German steels (K-2).

Figure 4.5-5 Distribution of nmm values for cold-rolled steels. Both parameters are determined by

the processing used to produce the steel and are not related to each other (K-2).

Figure 4.5-6 The combined relationship of FLD0, sheet thickness, and work hardening exponent (n) for low carbon steel (K-29).

Figure 4.5-7 Effect of yield strength, blankholder, and restrike on the springback on two automotive components (Y-1).

Figure 4.5-8 Strain distributions for four HSLA steels measured on the wing radius of an automotive bumper (A-12, N-7, K-33).

Figure 4.5-9 Relationship between yield strength of the steel and the equivalent strain in the center of the panel (S-7).

Figure 4.5-10 Formability balance for six steels shows wide formability differences depending on

mode of deformation (S-18).

5.1 Introduction
Lubricants in sheet metal forming have many purposes. Some are applied at the steel mill at the time the steel is produced to prevent rusting; they remain on the steel and become a primary aid to forming. Some are applied in the stamping plant after blanking. Others are contained in blank washers which serve the dual purpose of blank cleaning and lubricant application. Finally, a few may be selectively applied within the press to assist metal flow in a critical area of the stamping. The list of required lubricant characteristics usually is long. Cleanability, compatibility (with everything from phosphate treatment to adhesives), cost effectiveness, storageability, weldability, toxicity, solubility, and even formability are included; the list increases yearly. This review, however, will be restricted to formability characteristics. The other characteristics become separate items for discussion by themselves. The theory of lubrication has been covered in detail (S-11, S-12, W-23, B-13, R-1, K-3, F-5). An especially important work is TRIBOLOGY IN METALWORKING FRICTION, LUBRICATION, AND WEAR by John Schey (S-11). Therefore, no attempt will be made to repeat the various theories of lubrication. Finally, this review encompasses a rather broad view of the definition of lubricant. In addition to the traditional lubricants in liquid, paste, or solid form applied to the substrate metal, this review treats metallic coatings, such as anticorrosion coatings, as lubricants. For example, galvanized steels are analyzed in terms of a) the substrate steel with its primary control over the formability of the sheet metal and b) the interface composed of the zinc coating, the lubricant, and any die coatings. In a like manner, tin is considered part of the lubrication system in the forming of drawn and ironed beverage containers and paint becomes part of the lubrication system for prepainted stampings found in several industries. In terms of formability, lubricants have two primary functions. The first function deals with the control of sheet metal movement. This movement may take place from the binder area into the cavity of the die or may occur over the radius of the punch. The lubricant may encourage metal flow through a low coefficient of friction or restrict metal flow through a high coefficient. The key is a reproducible behavior which will consistently duplicate exactly the metal flow intended by the die designer.

The second function of a lubricant is prevention of scoring and galling. Here the lubricant must maintain sufficient isolation between the work piece (sheet metal) and the tool to prevent metal accumulation on the tooling (galling) which eventually leads to metal plowing (scoring). In a similar manner, wear of the tooling must either be avoided or reduced to a tolerable minimum.

5.2 Control of Metal Flow


In sheet metal forming, attempts are made to characterize the ability of a lubricant to control metal flow. This empirical characterization is performed without any attempt at understanding the reason or theoretical basis for the results. This characterization can be divided into four main areas: 1. Measurement of coefficient of friction 2. Measurement of metal flow during a simulative test 3. Measurement of sheet metal behavior during actual production 4. Characterization of the sheet metal surface Each area will be discussed in turn.

5.2.1 COEFFICIENT OF FRICTION


The first coefficient of friction measurements were conducted in a manner typical of a first course in Physics (S-1, W-30). A strip of metal was pulled between two flat platens whose surfaces are wider than the strip (Figure 5-1a). The pulling load resulting form the applied normal load was used to calculate the coefficient of friction. The test was simple. No deformation of the substrate occurred. Different combinations of substrate surface, substrate coatings, interface lubricants, tool coatings, and tool substrates could be readily evaluated. However, a sharp entry angle could not be tolerated because the lubricant would be scraped off. Any modification of the entry zone by increasing the radius would modify the type and degree of film developed in the interface. To overcome this die entry problem, a number of modified tests were developed. These modified tests included two cylinders (Figure 5-1b) used by Ike et al (I-1) or a combination of roller and platen (Figure 51c). A review of these tests has been provided by Miyauchi (M-32). These test fixtures allow for rapid evaluation of die composition and surface effects by interchangeable inserts. Observations of the recorded pulling load versus strip travel can lead to some general observations about the lubricants (G-18, A-9). A steady draw force indicates a good lubricant. An increasing force indicates gradual lubricant breakdown. A decreasing force suggests activation of additives or a change in the lubricant and/or surface behavior. A sudden drop at the beginning of the test shows a high static coefficient of friction compared to the dynamic value; this has been correlated to a tendency toward galling (G-18, A-9). Additionally, tests frequently are conducted with incrementally increasing force to indicate critical die pressures. Recently, Nine (N-14, N-15) has argued that the substrate in production stampings is deformed as the sheet metal stamping is created. One of the most severe deformations zones is the bending and unbending experienced when the sheet metal is drawn over a draw bead. Depending on the direction of the bending (positive or negative strain) and whether the sheet metal is in free space or in contact with the punch during the deformation, the sheet surface may roughen or become smoothed. This modification of the sheet surface can change the response of the interface lubricant. To measure the actual coefficient of friction during draw bead deformation, the flat platens were replaced by actual draw beads (Figure 5-2). The pulling load through the fixed bead set, however, is composed of two components the load to overcome friction (the desired measurement) and the load to deform the substrate during its journey through the bead set. To obtain only the friction load, a second test is

required. Here a strip of the same material is pulled through a frictionless bead set roller beads to obtain the load required to deform the substrate as it traverses the bead set. Subtracting the deformation load (roller bead) from the total load (fixed bead) yields the frictional load. From this and the normal a coefficient of friction can be obtained. Another common mode of deformation is observed at the die radius (Figure 5-3). Here a single bending and unbending is observed over a ninety-degree radius (L-13). To simulate die conditions more realistically, Woska (W-31) applie a blankholder pressure. Finally, Doege and Witthuser (D-20) utilized a roller to separate the frictional force form the other forces (much like Nine above) and to permit calculation of the coefficient of friction. This test also is used by Stine (S-40), but grease forced into the interface under pressure is used instead of the roller. These last tests, therefore, require three measurements: the bending plus the blankholder force, the frictionless bending plus the blankholder force, and only the blankholder force. Yet another mode of deformation the punch radius was studied by Duncan (D-22). Here, very little relative motion occurs between the tool and the sheet metal. A fixture to duplicate this mode of deformation is shown in Figure 5-4. As the strip is deformed, elongations in the vertical and horizontal legs are measured. By comparing the two elongations on the stress-strain curve, the amount of stress transferred around the pin, and therefore the coefficient of friction, can be calculated. Comparative studies using the different coefficient of friction tests described above have shown different coefficients of friction for a given material/lubricant combination when tested with a constant bead/die/punch radius (S-40, K-16, K-17). This result is not surprising. Mathematical models have been shown to be more effective in predicting actual measured strain distributions when the various coefficients of friction for the various forming modes are entered into the equation as opposed to a single value (S40). In addition, different lubricants respond differently when applied to different locations in the stamping. The problem becomes more complex when these tests are conducted on galvanized steels, because certain deformation modes are sensitive to the specific zinc coating/lubricant combinations, while other combinations are not (K-16, B-16, B-17, M-25). A numerical coefficient of friction is obtained from each of the above tests. However, Schey (S-11) and others warn that these tests represent only one set of test conditions. For example, test speed is one variable. In the draw bead simulator, Keeler has shown that some metal/lubricant combinations are speed sensitive, while others are not (K-17). While press shop personnel catalog forming operations in terms of inches per minute some at rather high speeds all sheet metal blanks must pass through an infinite number of speed regimes from zero to maximum forming speed (albeit for an extremely short time period) on their way from no motion at the start of the forming operation to maximum deformation rate. Other variables include increased ambient die temperatures form extended operation, lubricant buildup, lubricant breakdown, metal transfer to dies, interface pressure changes due to metal thickness changes, etc. Add these to test procedure variations, such as specimen cleaning, lubricant application, and test repeatability, and the resulting numerical coefficients of friction must be interpreted with respect to the variability of test parameters and the particular combination of parameters associated with the value quoted. Perhaps the best use of the numerical coefficients of friction obtained from the various tests described above is in the mathematical simulation or finite element analysis models. At least here the resulting changes in metal flow, strain distribution, loading/unloading, and other forming parameters can be determined through the various computations which tend to refine the impact of the raw coefficients of friction.

5.2.2 FLOW DURING A SIMULATIVE TEST


The coefficient of friction measurements described in the preceding section require careful attention to laboratory procedures, such as specimen preparation, repetitive tests, and numerical calculation. To short

circuit some of these problems, numerous laboratory tests have evolved in an attempt to stimulate common metalforming operations. Two common ones are the dome test and the cup draw test (M-24). Based on the differences in coefficient of friction described above, it is understandable that wide variations in test results are obtained. The situation is best described by the following segment from Schey (S-11): There are those who deny the validity of any simulation. There is, of course, some justification for such pessimism, as shown by the general lack of correlation even in such seemingly simple cases as evaluation of lubricants for sheet metalworking. Thus, Gibson et al (G-18) found no correlation between strip drawing and a cupping (deep drawing) test, nor between laboratory tests and production performance, particularly when the evaluation is based on friction alone. If, however, the trends in force were observed, correlation became much better, indicating lubricant breakdown is of much more decisive influence than the magnitude of initial friction. ..trends in forces, torques, etc. are very sensitive to changes in lubrication mechanisms and that it is therefore essential for the test run to be long enough to establish steady-state conditions. Details of the surface profile, which determine the load-bearing area and the fullness of the profile (which in turn controls the amount of lubrication entrapped). interact with interface pressure to change friction form low to high and vice versa.. Therefore, more important than the value of friction is whether the lubricant can resist breakdown, metal transfer, and galling. For this, testing at elevated temperatures is essential. Results are more reproducible and transferable when plastic deformation is induced in the course of testing. Thus, Ebben (E-3) reported on the evaluation of some 20 lubricants; correlation between a laboratory ironing test and plant performance was satisfactory. It should be amply evident form the above that no single test, or even an extensive battery of tests, can be used to rank the behavior of a lubricant in a complex sheet metal forming operation with all its interactions unless performed for an extended period such that elevated temperature and other steadystate conditions are reached. Evaluation of the lubricant performance with regard to metal flow should incorporate the latest evaluation techniques. For example, percent breakage is inadequate to evaluate forming severity. Instead, a technique such as Circle Grid Analysis (K-24, A-12, D-18) is a useful evaluation tool. The evaluation of lubricants, in terms of the other extremes loose metal, low spots, buckles, waves, springback, and other elastic/plastic behavior is less formalized. The ultrasonic thickness gage (K-11) can be utilized to detect small tensile versus small compressive strain states. The ultimate test for these defects, however, may be the human eye in the green room, where special lighting and highlight oils are use to improve detection of visual surface defects.

5.3 Sheet Metal Behavior During Production


Sheet metal behavior during actual press production is difficult to define and measure. As previously mentioned production is a complex interaction of many variables so that the effect of one variable is difficult to observe, much less measure. Therefore, most information is restricted to an obvious lubrication effect - scoring and galling. During deformation of the sheet metal over a tool, contact occurs only at the peaks of the sheet and the tool surfaces. Depending on the morphology of the surfaces, high contact pressures can be generated which cause microwelding of the asperities. Upon surface sliding, these welds are sheared. The resulting debris may then adhere to the tool surface and scratch the workpiece (R-16, H-28). This phenomenon has been called scoring, galling, and metal pickup. Even with a lubricant, plastic deformation of the surface asperities may lead to prow-like debris (R-11). Many investigations have been conducted to define the effect of sheet metal roughness on galling (M-32, L-13, H-28, R-11, F-11, K-8, G-19, G-25, Z-1, O-1, I-5, E-6, H-30, H-36). These studies have shown that non-galling surfaces generally had an arithmetical average roughness of Ra = 0.7-1.5 micrometer and a peak density of 3.7 5.7 peaks/mm. Less roughness and greater peak density were shown to be

detrimental because the surface area became too large compared to the valleys. The debris generated could not become trapped in the valleys to be removed from the deformation zone and the lubricant therefore had only minor reservoirs. Greater roughness and smaller peak density leads to too small a contact area between the sheet and tool, thereby creating a very high surface pressure. The influence of the sheet and tool material, lubricant, surface pressure, sliding length, surface roughness of the tool, sliding viscosity, and surface temperature on the friction and galling behavior in press working was investigated by Kumpunamnen (K-58) using a bending under tension type strip drawing tester. He concluded that good lubricants usually make the friction coefficient decrease with increasing surfce pressure; in this case, the coefficient of friction does not depend on the sliding length. With poor lubricants, however, the coefficient of friction increases as a function of both surface pressure and sliding length, and galling may occur. In contrast, many of the materials tested could be drawn over the steel tool bead without lubrication. Increases in sliding velocity always decreased the friction and often prevented galling. The influence of temperature, of course, depends on the chemical constituents of the lubricant. Kumpulainens study (K-58) also showed that the rougher beads increased the coefficient of friction resulting in an increased incidence of galling. If, however, the sliding direction and tool grinding direction are transverse to each other, the coefficient of friction decreases with increasing bead roughness: the transverse troughs act as reservoirs for both lubricant and debris. This study also showed that the bearing area provided useful qualitative information about the influence of various factors on the contact ratio and friction. Another study (M-27) showed that very clean surfaces lead to susceptibility to galling during sheet metal forming. The Japanese Deep Drawing Research Group studied galling (J-7) and concluded: a. Galling is strongly affected by contact pressure and sliding distance. Therefore, blankholder pressure, sheet thickness, and panel depth should be minimized. b. Blankholder force should be distributed as widely and uniformly as possible. c. Surface roughness of the die should be minimized, especially in the case of tool steels. d. The surface roughness of steel sheets does not seem to be a predominant factor within the present roughness range in commercial steel sheets. e. Die hardness does not play a major role in galling. However, the die hardness should be at least two or three times the sheet hardness to protect the die surface. f. The galling resistance is affected by the metallurgical properties of the die material and is especially improved in nonferrous die materials (see K-52). g. Higher strength steels have decreased galling resistance because of the larger forming force and resulting contact pressure. h. Galling resistance is improved by lubricants with higher viscosity. Sold lubricants are especially effective.

5.4 Standard Tests for Evaluating Formability Effects


ASTM Committee D-2 on Petroleum Products and Lubricants recognized the multifaceted problem of evaluating lubricants. Instead of a single test, the current procedure has four tests, each evaluating a different mode of deformation (A-27). These four tests are: 1) sliding strip test with flat dies, 2) sliding strip test with beaded dies, 3) biaxial stretch cup test, and 4) deep draw cup test. Relative to the Forming Limit Diagram, the first two tests are located at plane strain, the third is in the right hand side of the diagram,

and the fourth is in the left hand side of the diagram. Each test will yield a different lubricant ranking unique to its deformation mode, the substrate, and the lubricant.

5.5 Characterization of Sheet Metal Surfaces


More fundamental than simulative tests in the laboratory are attempts to mathematically describe the surface topography. A number of studies have attempted to define new parameters to define surface topography (A-5, A-12, B-11, V-6, L-12, M-3, D-9, A-17). One promising descriptor is the percent bearing area curve (W-12). The surface roughness of the sheet changes during deformation, depending on the deformation mode and type of interface contact with the tool (N-14, T-6). Thus, metal contacting a punch may have been so dramatically deformed being pulled over a die radius that description of the initial (asreceived and undeformed) surface topography may bear little resemblance to the sheet metal surface seen by the deforming punch. Unfortunately, the human eye seems to be the best device currently available for detecting differences in surface topography. In the future, non-contact devices which can detect and characterize surface topography features identifiable to the human eye will be required to differentiate between different products.

5.6 Summary
Lubrication serves two functions in sheet metal forming: modification of friction to control sheet metal movement and prevention of scoring and galling. Friction occurs between the metal substrate (workpiece) and the deforming tool. In this context, the lubricant includes not only liquids or solids introduced into the interface, but also metallic precoatings of the steel. Studies have shown that a different coefficient of friction is associated with each combination of sheet steel surface, metallic precoat (zinc, tin, paint, etc.), mill applied lubricant, press shop applied lubricant, and tool surface. In addition, for a given combination, the coefficient of friction is a function of the specific mode of deformation to which the sheet steel is subjected. Thus, many opportunities exist to tailor this combination to provide the desired metal flow in different zones of the stamping. The correct combination usually is selected by trial and error. The proliferation of surface, coating, and lubrication combinations which provides extensive press shop options also make s prediction of specific coefficients of friction difficult for each combination. These coefficients of friction are required, however, for accurate mathematical modeling of formability and other sheet metal forming predictive techniques. Determination of these coefficients of friction by laboratory tests depends on how accurately the laboratory tests simulate the actual forming operation, including speed of forming, metal flow rate, interface pressure, and interface temperature. Rules of thumb generally are found in the literature relating to techniques for avoiding the onset of falling and scoring.

Figure 5-1

Figure 5-2

Figure 5-3 A draw bending friction test fixture simulates the deformation created when a strip is pulled over a die radius with constraint from the binder pressure.

Figure 5-4 Frictional conditions around the radius of a punch are simulated by this test fixture by

Duncan (D-22).

6.1 Introduction
The press is one of the major components of the forming system. Press shop experiences have shown that a specific material/lubricant/die combination can produce satisfactory stampings in one press line and be incapable of producing satisfactory stampings in another press line. This is especially true for stampings which lack sufficient safety margin and are balanced at the onset of failure. Likewise, the ability to produce quality stampings without loose metal, buckles, waves, low spots, and other surface defects depends on the characteristics of the press in which the stamping is produced. Some presses have very accurate alignment and tight guidance systems. Other presses have loose guidance systems and the punch to die alignment, as well as blankholder reproducibility, changes from one stamping to the next. The sensitivity of the forming system to the characteristics of the press and the prevailing adjustments is confirmed by traditional press shop practices. When stamping quality needs to be improved either with respect to failure or surface appearance one of the first adjustments to be made is to the press. These adjustments include position of the holddown or binder ram, position of the main ram, ram speed, line pressure on cylinders, etc. This document is a review of sheet metal formability. Acknowledging the role of the presses as one component of the forming system is the purpose of this short section. The details of press construction,

design features, maintenance, etc. are beyond the scope of this review. The reader is directed to reference books for this information, such as the textbook by Eary (E-2). The important point is that variability of the press either intentinal or undetected contributes to the variability of the system.

6.2 Press Characteristics


The two main press drives are mechanical and hydraulic. Many discussions have been held about the advantages of both without any real definitive conclusions being documented. One of the main differences between the two types is the load/stroke characteristics. The hydraulic press has a constant load capacity throughout the entire stroke, whereas the mechanical drive reaches maximum load capacity only at the bottom of the stroke. This becomes more important for pure stretch forming operations which require maximum blank clamping loads to be exerted at ram positions far removed from the bottom dead center and high up on the stroke curve where load capacity is low for mechanical presses. Another difference is the length of the stroke. A hydraulic press can be set easily at any stroke within the limits of the hydraulic cylinder travel. The ram stroke on the mechanical press is fixed. A third difference is the speed curves. The hydraulic press can operate either at a constant ram speed or any programmed speed cycle desired. The speed curve of the mechanical press is neither constant nor programmable, but is a fixed sine wave. The speed ranges from zero at the beginning and end of the stroke to maximum speed at the halfway point. This particular speed curve may not be ideal for some critical stampings. However, minimum information on this effect appears in the literature. In general, the differences between press characteristics can easily be defined. The problem remains to conduct some definitive experiments which will document the effect of these differences on sheet steel formability.

6.3 Press Instrumentation


Instrumentation of the press is a key area for advancement of sheet metal formability especially in terms of efficiency of die change and reproducibility. If the successful forming of metal depends on the relative flow of metal from the binder, then binder loading now becomes an important factor to both measure and control. Load monitors on all four corners of the inner and outer rams, plus digital position indicators for the same locations, become important and are being installed on many presses. The primary use of such instrumentation is initial setting of the press parameters when a die change has occurred. The previous successful settings are used to reset the press when the die is reinstalled. A secondary use is to monitor the process (forming operation) during an extended run. If the load readings are changing, this could indicate that the forming process itself is changing. Changing load readings suggest that the quality of the stamping is changing. Currently, a key topic is rapid die changes. While the dies can be physically changed in a rather short time, the problem lies in obtaining good quality stampings within minutes after the completion of the physical die change. Press instrumentation can assist in the adjustment of the forming system to the ideal state. The main ram and holddown ram positions are set on the digital indicator to either the previous setting or a specified ideal setting established during die tryout. During the first piece run, the load indicators should reach values in the previously determined acceptable range. If the load readings are different, it means that the forming system is different from the previous run or the ideal condition depending on which settings are used. Further investigation is required to determine the cause of the variation, such as material out of specification, change in lubrication, misadjustment of the dies, etc.

6.4 Summary

Presses have always been one component of the Forming System. However, traditionally they have been considered as a constant factor and generally ignored. However, system reproducibility demands that the conditions of the press be factored into the system equation. Measurement and control of press characteristics become important topics in the study of sheet metal formability. In a like manner, the specific adjustments to the press must be measured, documented, and reproduced for each set of tooling. Two specialized forming operations are reviewed in this section. The first is stretch forming, which is rapidly gaining acceptance for many outer body panels. The other is roll forming. Long used by other industries, roll forming is gaining additional acceptance by the automotive industry.

7.1 Stretch Forming


7.1.1 INTRODUCTION
Stretch forming is the simplest of all the sheet metal forming operations to analyze and understand. The ability of sheet metal to be stretched can readily be duplicated in the laboratory. The properties of work hardening, strain rate hardening, instability, and fracture have been studied extensively and defined (K24, G-10, K-28). The stress rate is considered to be plane stress with the only important stresses acting in the plane of the sheet. The interface stress, while important for friction behavior, is small relative to the yield stresses. The stretch forming mode is simplified because all the metal flow from the binder area is eliminated. This greatly reduces the complexity of the deformation. In addition, a smaller blank size is utilized because additional surface area is generated by the stretching operation. In contrast, metal flowing into the die area from the binder area can thicken and reduce available surface area due to the compressive stresses. Stretch forming has excellent potential as the primary mode of deformation to be used by press shops in the future. The advantages of stretch forming will be developed in detail later in this report. However, the list includes: ease of analysis, no metal flow from the binder, reduced blank size, tight panels, reduced sensitivity to metal variability, and increased flow stress in the formed panel. The minimization of residual stress gradients reduces springback and improves final dimensional accuracy to part print. Compared to aluminum and other metals, sheet steel is well suited to the stretch forming mode of deformation (H-16, K-9). Most important is the positive strain rate hardening exponent, which is related to post-uniform elongation. This is reflected in the high level of the Forming Limit Diagram. Also important are the high levels of the work hardening and the resistance to sheet thinning (as reflected by the normal plastic anisotropy) of steel. Several laboratory advances have enhanced the understanding of sheet metal stretching and opened the way for more effective utilization of stretching as the primary mode of deformation in the future. One is the computer simulation of sheet metal forming (K-48, A-3, B-15, W-10). All the interactions of the stretching deformation can be simulated and the ideal combination selected. The advantages of stretch forming then become obvious.

7.1.2 BASIC DEFINITIONS


7.1.2.1 Stretch Versus Draw
Stretching and cup (deep) drawing are basically opposite forming processes. (See discussion in Section 2.3). The difference is highlighted in Figure 7.1-1.

Stretch forming is deformation over the punch. This deformation mode does not allow any metal flow from the binder (metal under the blankholder). Only metal initially within the die opening undergoes strain. This deformation mode results in a decrease of metal thickness. In deep drawing, metal under the punch (initially within the die opening) undergoes no strain. All deformation takes place in the sheet metal under the blankholder as it moves towards the die opening. An additional requirement is that both the blank and die opening be circular. Therefore, as metal moves towards the die ring, it is compressed in the circumferential direction (the circumference is constantly decreasing for all elements) and elongated in the radial direction. This results in an increase in metal thickness. A third deformation mode found in most stampings is termed bend-and-straighten (Figure 7.1-2). This is similar to deep drawing except that the die radius is a straight line. Thus, an element is bent as it conforms to the die radius and then is straightened as it enters the wall of stamping. Because the die radius is a straight line, no decrease in the circumferential direction takes place. The metal exiting into the die wall ideally has no net change of shape, although metal work hardening has taken place. The breakdown of a complex stamping into component forming modes (K-26, A-12) will serve to illustrate the difference better (Figure 7.1-3). The straight-line segments of the walls are formed by bend-andstraighten type of deformation. Only the corners are formed by deep drawing. The metal below the restrike line (initial location of the blank at the die radius) is stretch formed. In the stamping illustrated in Figure 7.1-3, only a small portion of the part is generated by deep drawing. In some parts, the entire stamping is created by only stretch forming or stretch forming plus bend-and-straighten.

7.1.2.2 Stress State Versus Strain State


One problem in identifying modes of deformation is that identical strain states can be created by different stress states. This is illustrated in Figure 7.1-4. An identical biaxial strain state can be generated (Figure 7.1-4a) by either a compressive stress through the sheet thickness or by equal biaxial tensile stresses in the plane of the sheet. In this case, the thickness strain is the same for both deformation conditions. Only the latter deformation condition is stretch forming. Here the stress state must be know to identify whether stretch forming has taken place. Another set of common stress conditions illustrated in Figure 7.1-4b creates the same pair of surface strains. However, in this case the compressive stresses cause an increase in sheet thickness. The tensile stress causes a reduction in sheet thickness. Again only the latter condition is considered to be stretch forming. Compressive stresses in the plane of the sheet are not generated in stretch forming operations. The definition of stretch forming present in the first section (deformation within the die opening) now can be generalized an redefined in terms of stress state. Stretch forming is uniform deformation through the sheet thickness where the major stress is positive and the minor stress is either zero or positive. The requirement of uniform deformation through the sheet thickness eliminates bending from this definition. This definition encompasses a wide variety of stretch forming deformation states; several possibilities are illustrated in Figure 7.1-5 for an isotropic metal.

7.1.3 MODES OF STETCH FORMING


Many different forms of stretch forming have been used over the years to produce a variety of sheet metal stampings. Except for the stress definition of stretch forming presented above, little apparent commonality exists among the different production processes using some form of stretch forming to produce stampings. Therefore, a tabular method is used here in an attempt to provide at least some form of grouping to highlight the types of operations utilizing stretch forming.

Categorizing production processes according to this tabular format is complicated in that some processes combine more than one of the fundamental modes of stretch forming. These more complex production processes therefore are not listed in the Table I. TABLE 1 - TYPES OF STRETCH FORMING
SHEET FORMING

Concept Pre-stretch Locked Stretch

Method Clamp and Pull Before Form Lock Beads During Form Cyl. blank Lock and Pull After Form
EDGE FORMING

Post-Stretch

Example Cyril Bath (B-7) Automotive Stretch Dies Fluid Forming Wallace Expander (R-17) Shapeset (A-32) Frame Rail
Built-in Threads

Stretch Flange
Hole Expansion (Extrusion)

Wiping
Pierce and Extrude

7.1.4 PRODUCTION PROCESSES


As seen in Table I, a variety of stretch forming processes is available, either singularly or in combination. To better understand the stretch forming modes, each of the production processes will be described.

7.1.4.1 Cyril Bath Stretch Former


The Cyril is a machine used to form panels much like a standard closed-die system (B-7). A male punch and female die are used to generate the final forming. However, a special clamping device replaces the binder ring. In operation, the gripper jaws clamp the blank and prestretch it a fixed amount. This accomplishes four things. First, a smaller blank can be used, reducing blank cost. Second, the stretching of the blank removes loose metal. Third, all areas of the blank receive some preset degree of cold work, thereby strengthening the otherwise unworked areas. Fourth, a more uniform strain distribution is generated, reducing the severity of the stamping. The last benefit above requires more explanation. In a normal forming operation, deformation begins at a location of high stress. When the level of stress reaches yielding (point A in Figure 7.1-6) those elements subjected to the high stress yield and work harden. However, adjoining elements are at lower levels of stress. Some could be as low as point B. Therefore, point A has to undergo extensive straining and work hardening (to point C) until the stress at B climbs sufficiently to reach yielding. The resultant strain distribution would be extremely non-uniform. In many forming operations, element B never reaches yielding. This is especially true around sharp radii and under flat punch areas. If the blank was prestretched to yielding and then pulled over the male punch without unloading the stress, a change in behavior would be observed. Point A would undergo an additional increment of strain and work harden. However, the increment of work hardening would be small before the adjacent element (shown as point E in Figure 7.1-6) would begin to undergo deformation because it was already preloaded to the same yield stress. Thus, the strain distribution would be much more uniform.

The Cyril Bath Stretch Former therefore produces tighter stampings from smaller blanks. In the press shop, a tight panel is one which conforms to the design shape without buckles, loose metal, low spots or depressions, lack of rigidity, or other specified defects. The resultant stampings are less severe because of the improved strain distribution. Therefore, this forming technique is more applicable to less formable metals. Finally, shallow panels can be more easily formed by this technique because they do not have a vertical draw wall which is difficult to form. Consequently, most Cyril Bath machines are found in the aircraft/space industry. Only very limited trials with these machines have been conducted by the automotive industry.

7.1.4.2 Stretch-Draw
The stretch-draw system of die design is a modification of the conventional toggle draw die system. The two differences are that the die set is inverted and that a double acting toggle press is not required. In stretch-draw, the male punch is in the bottom position. As such it is attached to the bed of the press and does not move. The die cavity and ring, now in the top position, are lowered by the action of the press. The lower ring is not fixed but is located on an air cushion or a system of nitrogen die cylinders. This removes the need for the outer slide of the press. The first action is the same as the conventional toggle draw die the binder shape is set with the rings. This is accomplished by the upper die portion moving down. During the second action, the upper die cavity/ring moves down over the punch. This is identical to the punch moving into a stationary die cavity. The lower die ring moves downward on the air cushion or nitrogen die cylinders while maintaining the preset force. These cylinders are nitrogen filled and act much like shock absorbers. Placed under the lower die ring, they exert a constant upward force while collapsing to allow the die rings to move downward. One advantage of the stretch-draw process is that a single action press can be used to provide double action performance. This is especially advantageous when an older press can be utilized in this manner. Alternatively, the technique can be utilized in a transfer press which can not provide independent double action for each station. The stretch-draw die system does not, however, eliminate the problem of controlling metal flow into the die cavity from the binder area.

7.1.4.3 Inverted Toggle Draw With Floating Binder


As the name suggests, this die set is similar to the conventional toggle die set. However, there are two differences. First, the die set is inverted with the punch fastened to the bed of the press and the die cavity moved by the main press ram. However, unlike the conventional toggle draw or the stretch-draw dies, this die configuration has both the upper and lower binder rings independent of the punch/cavity system. This allows an extra degree of freedom in positioning the blank; this extra action can be used to stretch the blank over the lower male punch. The first action is to set the blank in the binder. This setting can be done with the blank supported by the male punch. This controls blank sag and blank shape when the binder rings are closed. The second action is the outer slide pushing down on the upper binder ring. The lower binder ring moves down in a controlled manner due to the action of the pressure cylinders which maintain the binder locking action while collapsing downward. This causes the general male shape to be stretch formed into the blank which exerts a general tensile pull over the entire blank. This action is absent in both the conventional die system and the stretch-draw system. The third action is the upper or female die closing on the blank. This

not only imparts the final panel shape, but the outer edges of the female cavity act as secondary stretch points to provide an additional stretching action.

7.1.4.4 Wallace Expander


The Wallace Expander is used by the appliance industry to produce washer/dryer tubs, outer wrappers, and inner liners (R-17). All these items are contiguous, four-sided stampings. The expander system begins with a cylinder of sheet metal formed by rolling a tube from a blank and welding the junction line. The cylinder is then placed over a segmented punch. This punch can be segments of an arc (eight or more) to generate an expanded cylindrical drum for washers or dryers. The segments can also be foursided flat punches to form four-sided wrappers and inner liners. Inside the segments is a conical wedge. After the cylinder is placed over the punch, the wedge is driven upward. This forces the segments outward. After contact with the cylinder, continued expansion places the cylinder in tension. In some cases embossments are subsequently formed in the sidewall of the cylinder. For this, a mating set of male and female die inserts are placed in the punch and cavity walls. Because the cylinder is both in tension and acts like an infinite blank, the deformation mode of these subareas is by embossing, whereby deformation is confined to the punch and die radii. Applicable forming limit rule of thumb must be observed, such as the height of the embossment can be no greater than 80 percent of the sum of the punch and die radii for steel. The Wallace Expander concept fills a unique need for the appliance industry. Here the stretched formed stampings have excellent shape without loose metal. All the surface area has been subjected to a tensile strain. This not only work hardens the steel but also causes sufficient deformation to trigger age hardening in rimmed steel.

7.1.4.5 Post Stretch of Channels


A specific forming operation channel forming has been the focus of much research in stretch forming. The channels are formed by the traditional forming processes of bend-and-straighten. This primary forming operation usually results in springback, which takes the form of sidewall curl show in Figure 7.1-7. post stretch forming is used to eliminate this curl. Research in this area has been carried out by Ford (D-3), GM (A-32), and other companies. GM has a formal production practice called Shapeset (A-32). All these techniques involve some form of post stretch. Here the channel flange is locked near the end of the stroke while the punch continues down to impart tensile deformation in the channel wall. With proper die design conditions, any degree of curl can be obtained even negative curl as shown in Figure 7.1-7. Early work by GM-APMES (J-5) and Duncan (D-23) showed interesting results (Figure 7.1-8). Springback in channel forming was changed by the amount of post-stretch forming, where the channel flange was locked before the end of the stroke. Post-stretching reduced springback. The higher the strength of steel, the greater the post stretch required before the minimization process begins. Thus, for some level of poststretch found in most channels (Line A in Figure 7.1-8) the amount of springback highly depended on sheet strength. Thus, to control springback in this case sidewall curl a narrow range of yield strengths is required. The important aspect of the work was that beyond a certain minimum amount of post stretch (shown as line B in Figure 7l1-8) all the steels reached the same level of springback (level C in Figure 7.1-8). Design considerations such as this are important for dual metal dies, such as low/high strength steel and steel/aluminum combinations. A similar type of post stretch can be found in restrike operations. The very nature of the restrike is to sharpen radii, impart details, and change configurations; this incorporates an element of stretch. The amount of stretch often exceeds the capacity of the sheet metal and tearing results in many attempts to restrike.

7.1.4.6 Specialized Stretch Operations


A series of specialized forming operations are pure stretch forming by their design. They include: explosive forming magnetic forming electro-spark discharge hydroforming. In all cases a blank is securely clamped to the binder ring. In the case of the first three listed above, a shapeless punch is used to push the metal into the cavity. The nature of the shapeless punch (some type of fluid) requires a complete locking of the periphery to insure a seal. This creates the stretch forming mode. In the case of hydroforming, the female die cavity is replaced by a rubber pad or by a rubber diaphragm over a cavity filled with oil.

7.1.5 MATERIAL PROPERTY INTERACTIONS


7.1.5.1 Forming Limit Diagram
The Forming Limit Diagram or FLD defines the maximum amount of deformation a sheet of metal can undergo before the onset of localized necking terminates useful deformation. This deformation is defined in terms of two principal strains called the major and minor strains. The Forming Limit Diagram (Figure 7.1-9) shows that the amount of major strain a sheet can under go is determined by the sign and magnitude of the minor strain. Early work on the FLD (K-12, K-21) related the left side of the curve to deep drawing and the right side of the curve to stretch forming. This description overlooks the fact that the left side of the curve also describes an important range of stretch forming. For example, the tensile test plots on the left side of the diagram because it has a negative minor strain (equal to half the true major strain for an isotropic metal). Therefore, the entire FLD can be used as one representation of the stretching limit of sheet metal. This can be more clearly visualized if the FLD is modified by plotting the thickness strain on the vertical axis (Figure 7.1-10). This shows a constant thickness strain as the forming limit for negative minor strains (K11). Therefore, my forming component such as pure shear which does not decrease the sheet thickness is a desirable component of any stretch forming operation. The importance of the FLD to stretch forming technology will be its importance to the part/die designer. Once the sheet metal grade and thickness are specified, the FLD is fixed. However, the strain ratio usually is a variable which can be controlled by proper design of both the part and the forming dies. Thus, the FLD not only indicates to the designer the current stamping severity but also identifies options which are available to modify the severity. The FLD defines conditions for the onset of localized necking (K-15). This is the concentrated band of thinning (neck) which precedes fracture. Thus, substantial deformation can occur within the neck prior to fracture. However, because the deformation is confined to the neck, no practical overall deformation of the part is accomplished during this localization to fracture. Thus, the FLD represents a practical limit to useful deformation.

7.1.5.2 Work Hardening Exponent


Stretch forming capability of a material is controlled primarily by the work hardening exponent, n, defined by the stress-strain equation:

= K where is the current true stress to continue deformation, K is a constant for the metal under test, is the true strain, and n is the work hardening exponent. The n value influences stretch forming in two ways. First, the height of the FLD, defined by FLD0, is a direct function of the n value (A-12). This relationship is shown by the plot in Figure 7.1-11, which indicates that the FLD increases for increasing values of the work hardening exponent. This relationship is valid only for high-strength steels with yield strengths greater than 50 ksi (345 MPa). For steels with yield strengths less than 50 ksi (345 MPa), the level of the FLD has been experimentally shown to be constant (K-29). Second, the n value is one of the factors which determines the uniformity of the strain distribution; the other important factors are lubrication, part/die design, and press adjustments (K-23). A high n value means greater work hardening of the metal. This means the metal will more uniformly distribute the strain in the presence of a stress gradient (H-22).

7.1.5.3 Strain Rate Hardening Exponent


The strain rate hardening exponent or m value is determined from the strain rate hardening equation: = Km where is the instantaneous flow stress, K is a material constant, is the strain rate, and m is the strain rate hardening exponent. As the strain rate increases, the flow stress increases. Thus strain rate hardening acts much like work hardening in distributing the stain more uniformly (C-8). An increase in the m value increases the postuniform strain capability. This is reflected in the level of the FLD which describes the sum of uniform and post uniform deformation. While m value is relatively small for a given metal, the effect is large. It accounts for the reduced level of the FLD (about one-half) and the reduced stretchability of 2036-T4 aluminum compared to a low carbon steel with the same value of n (H-19).

7.1.5.4 Anisotropy Ratio


The anisotropy ratio, r, is a measure of the directionality of the uniform straining in sheet metal. It is commonly defined as the ratio of the width strain to thickness strain when a strip of metal is elongated by a uniaxial stress. If the two strains are equal, the deformation is termed isotropic. If the r value is greater than one, the metal resists thinning. The r value has not been shown to affect the height of the FLD (B-3). It will, however, affect the strain path the metal will follow when subjected to a given stress state. The best example is the uniaxial tensile test (Figure 7.1-12). For a given major strain, a metal with a high r value resists thinning. Because of the constancy of volume, this reduction in thickness strain must be accompanied by an additional increment of width strain (negative minor strain in this case). Thus, the strain path deviates towards the left and climbs to a higher major strain before reaching the limiting strain curve. Another way of describing the effect of r value is based on the critical thickness strain level on the left side of the FLD (Figure 7.1-10). Therefore, any factor such as the r value which generates a reduction in thickness strain allows for greater deformation before reaching the critical level of thickness strain.

7.1.5.5 Yield Strength

The yield strength of a metal influences stretch forming in several indirect ways. (i) As yield strength increase above 50 ksi (345 MPa), the n value decreases proportionately (Figure 7.1-13). This reduced the level of the FLD and reduces the ability of the metal to distribute the strain uniformly in the presence of a stress gradient. (ii) The increased strength causes a greater loading on the binder. Therefore, the locking system (beads, etc.) must be more effective to withstand the increased radial forces. (iii) The increased flow stress of the metal increases the interface pressure between the punch and the sheet. This means that the lubricant must be more effective and may require an EP (extreme pressure) additive to provide comparable lubricity. Therefore, unlike deep drawing which is independent of yield stress of the metal, stretch forming is reduced as the strength of the metal is increased. Cold work is not an effective method to strengthen steel which is to be subsequently used for stretch forming. In Figure 7.1-13, steel strengthened by cold work has n values well below the curve. This curve (B-4) was obtained for steels strengthened by grain size control, solid solution hardening, and/or precipitation hardening. In fact, small amounts of temper passing can be quite detrimental to n values (C6). The same comments are applicable to multiple stage forming, where prior cold work can severely limit stretch forming capability. Thus, for sequential operations, stretch forming followed by compressive deformation is more effective than compressive deformation (which reduces the n value) followed by stretch forming.

7.1.5.6 Elongation
Uniform elongation is related to the n value of the material. Since n value is the more fundamental property of the metal, stretch forming is better related to n. In addition, uniform elongation can be difficult to define and measure from a tensile test. Total elongation is the summation of a metals capacity for uniform elongation, diffuse necking, and localized necking which eventually terminates in fracture. Uniform elongation is related to n value and diffuse necking is related to the m value. Beyond the onset of localized necking, useful deformation no longer occurs for stretch forming over a rigid punch. Therefore, a measure of n and m better reflects the stretch forming capacity of the metal than the total elongation. The problems with total elongation are compounded by its dependence on gage length over which the measurements are made.

7.1.6 SIMULATIVE LABORATORY TESTS


Both the producers and users of steel for stretch forming applications desire some form of quick test(s) to prequalify the steel for its intended applications. Several options are therefore available. Specific formability evaluation of sheet metal currently is divided into two components. The first is the basic formability of the substrate itself. The second is the surface topography of the sheet metal and its interaction with the lubricant and tool surfaces. Therefore, two common methods of evaluating stretchability of sheet metal are (I) with the sheet surface in free space and (ii) with the sheet surface in contact with a lubricant/tool interface.

7.1.6.1 Tensile Tests


The most important substrate (bulk) properties related to stretch forming are the n and m values. For specific applications the r value and total elongation can also be important. All four of these properties can be obtained from the tensile test. By the design of the tensile test, the surface of the specimen is in free space and normal surface topography does not affect the results. Therefore, the tensile test is by far the best evaluation of substrate properties for formability. This is especially true for stretch properties where the property-stretchability interactions are well defined.

Arguments have been made that the tensile test is only a uniaxial stress test while the stress state in practical deformation in most automotive stampings is approximately plane strain. Deformation over the head of the punch often may be balanced biaxial. Biaxial stress tests are difficult, if not impossible, to conduct. In one attempt to evaluate the substrate properties under these stress states, hydraulic bulge test are conducted by some laboratories. The fluid punch has no interaction with the sheet surface. With extreme care, a biaxial stress strain curve can be obtained along with the attendant n and m values (J16). Experimental problems with hydraulic bulge tests, however, restrict this test to research laboratories at best. Included in this same category of tests are the Marciniak (M-11), reduced cross-section (A-35), and other specialized sheet metal formability tests which isolate the sheet surface from the tool/lubricant interface.

7.1.6.2 Punch Tests


Punch tests are an attempt to simulate the total forming system, which is basically the formability of the substrate modified by the interface interaction with the tool. Historically these punch tests have included the Olsen/Erichsen dome tests, the Swift round-bottom punch, Fukui, and other simulative laboratory tests. The latest punch test to emerge is the Limiting Dome Height Test or LDH test (G-9, A-31, M-23). Problems with the test procedure (M-23) notwithstanding, the use of the LDH and all other simulative punch tests raises a major issue about correlation. Different locations in a single stamping, much less different stampings, all have different combinations of stress state, interface pressure, lubricant/sheet surface interaction, speed, temperature, prior strain history, and all the other active variables. Even the most complex of computer forming simulation programs can not predict the resulting strain distribution and provide an estimation of the stamping severity. The problem is compounded when coated steels multiply the possible interactions among tool surface, sheet surface, and interface lubrication. How then can a single test of sheet metal correlate well with all stampings? Similarly, laboratory simulation of stretch forming of a blanked edge raises difficult problems to be overcome. Here the hole expansion test (D-6, D-8) is commonly used as an evaluation test. Even this simple test is completely dependent on the quality of the initial hole, the shape of the punch, and other test factors. For the above reasons, punch simulation tests should be used with caution and preferably only when tensile test data are unavailable.

7.7.7 SUMMARY
1. An important element of many complex forming operations is stretch forming. Press shop terminology, however, usually categorizes this stretch mode as deep draw. 2. Stretch forming is defined by a stress state in the plane of the sheet metal composed of a positive (tensile) major stress and a positive or zero minor stress. Simply defining the strain state is insufficient to differentiate stretch forming from some compressive modes of deformation. 3. Stretch forming always creates a reduction in sheet thickness. 4. The stretch forming limits for deformation over a rigid punch are defined by the onset of localized necking and are detailed by the Forming Limit Diagram. 5. The forming limits for stretch forming a blanked (free) edge are related to the tensile elongation of the substrate metal and the quality of the blanked edge.

6. The two most important metal characteristics for general stretchability are the work hardening exponent, n, and the strain rate hardening exponent, m. Both properties improve the distribution of stretch deformation by reducing the localization of strain during uniform straining and post-uniform straining, respectively. These properties allow good stretchability in steel. 7. The quality of sheet metal for stretch forming operations is primarily related to substrate formability. Surface topography and coatings interact with the lubricant and tool steel to modify the lubricity of the interface. 8. Stretch forming is well understood in terms of mathematical simulation and some success has been achieved for axisymmetrical stampings. 9. Most pure stretch forming operations, such as Cyril Bath Stretch Former and the Wallace Expander, are currently confined to non-automotive applications. Figure 7.1-1 Schematic showing the difference between (A) stretch forming with all deformation

within the die opening, and (B) deep drawing with no deformation within the die opening.

Figure 7.1-2 Schematic of the bend-and-straighten mode of deformation. Circle grids are

undeformed after the element enters the stamping wall.

Figure 7.1-3 A complex stamping is generated by many different forming modes. Only a small

section of most stampings is formed by a deep draw mode.

Figure 7.1-4 Schematic showing the range of possible stress states for stretch forming an

isotropic sheet metal.

Figure 7.1-5 Schematic showing the range of possible stress states for stretch forming an

isotropic sheet metal.

Figure 7.1-6 Stretching under pretension creates a more uniform distribution of strain because all elements of the blank are initially at the yield stress.

Figure 7.1-7 Sidewall curl caused by springback during forming is positive. By adding a post

stretch operation, the amount of curl can be reduced or even made negative (D-3).

Figure 7.1-8 Springback depends on the strength of steel until the post stretching exceeds a

minimum level indicated as B (G-6).

Figure 7.1-9 The Forming Limit Diagram indicates the proximity of a strain state to the onset of localized necking. Above the curve the conditions for localized necking are satisfied and the breakage is expected.

Figure 7.1-10 A modified Forming Limit Diagram in which the thickness strain replaces the major strain as the vertical axis. Note the constant thickness strain as the forming limit for negative

minor strains (K-11).

Figure 7.1-11 A nomograph showing the value of FLD0 as a function of the work hardening exponent, n, and teh sheet thickness (A-12).

Figure 7.1-12 The anisotropy value, r determines the strain path taken by a unixial tensile text.

Figure 7.1-13 A curve relating the work hardening exponent, n, to the yield (flow) stress of steel. Steels in the band were strengthened by grain size, precipitation hardening, solid solution hardening or a combination thereof. Steels strengthened by cold work have lower n values for equivalent strengths (B-4).

7.2 Contour Roll Forming

7.2.1 INTRODUCTION
Contour roll forming, or cold roll forming, is a process whereby a sheet or strip of metal is formed into a desired shape of uniform cross section by feeding the stock longitudinally through a series of roll stations equipped with contoured rolls or roller dies two or more per station. The forming is accomplished by progressively working the metal in each station until the finished shape is produced. The cold roll forming mill consists of a train of pairs of driven roller dies which progressively form the flat strip into the desired configuration. The key to successful forming of the desired shape is the correct determination of the number of pairs of rolls required to form the shape and the proper contour of each roll. An insufficient number of roll pairs will cause too severe a change in contour at each pair. Too many roll pairs may be beneficial to the forming sequence but add unnecessary cost to the operation. The number of pairs of rolls depends on the type of material (and any coating) being formed, the complexity of the shape being produced, and the particular design parameters of the mill being used (such as center distance between roll stands). A conventional roll forming mill may have as few as 4 or more than 30 pairs of roller dies mounted on individually driven horizontal shafts. Idler (undriven) sets of rolls can be placed between driven sets for additional shape control. The contour roll forming process is particularly suited to the production of long lengths of complex shapes held to close tolerances. Large quantities of these parts can be formed with a minimum of handling and manpower. The process can be continuous by coil feeding and exit cutting to length or batch operated by feeding individual sheets. Auxiliary operations, such as notching, slotting, punching, embossing, curving, and coiling can easily be combined with the contour roll forming to produce finished or semi-finished part shapes off the exit end of the roll forming mill. When the entry material is prepainted or otherwise coated, ready to ship parts can be shipped directly from the mill.

7.2.2 FORMING MODE


Contour roll forming is one of the few sheet metal forming processes which is confined to a single primary mode of deformation (Figure 7.2-1). Unlike most forming operations which have various combinations of stretching, radial drawing, bending, bending-and-straightening, stretch flanging, shrink flanging, and other forming modes, the contour roll forming process is nothing more than a carefully designed series of bends (A-14, S-29, G-43). The deformation mode in contour roll forming is bending both under the contoured rolls at each station and between each station. The metal thickness is not changed except for a slight thinning at the bend radii (G-43). Formulas for calculating the required initial width of the strip (discussed later) are simply a mathematical substitution for the manual method of determining blank width (R-14, J-15). The manual method consists of making a large-scale layout of the final part shape, dividing it into its component straight and curved segments, and totaling the developed width along the NEUTRAL AXIS. This is illustrated in Figure 7.2-2. Contour roll forming is performed on sheet/strip stock which varies from 0.005 to 0.75 inches (0.13 to 19 mm) in thickness and from 1/8 to more than 72 inches (3.18 to 1829 mm) in width. All metals that can be shaped by any of the common forming processes especially bending can be contour roll formed. The formability of the work metal controls the permissible speed of the roll forming and the degree or severity to which the metal can be formed. For example, the softest grade of aluminum can be contour roll formed at a speed 400 times faster than rolling titanium strip into a similar shape (A-14). In addition, the soft metals can involve bending through 180 degrees, while titanium might be limited to 90 degrees or less.

7.2.3 DESIGN CONSIDERATIONS

The cross sections which can be roll formed are as varied as the materials which can be formed into these cross sections. All texts on roll forming, however, agree that the most effective and trouble-free operations are those which are designed with the roll forming process in mind. A few simple rules should be followed (A-14). SYMMETRY - The ideal cross section has symmetry. Nonsymmetrical sections can be formed without difficulty, but lack the equal stresses imparted to each edge of the metal as it passes through the form rolls. CROSS SECTION DEPTH - Extreme depth in the cross section should be avoided. Roll forming produces more complex stresses than brake bending or other simple bending processes. In deep sections, the metal movement around the arc of the bend is much greater and the resulting edge stresses are greater. BEND RADII - Sharper radii can be obtain by roll forming than by other forming processes. However, the minimum bend radius is still a function of the ductility of the work metal. Bend radii equal to metal thickness should be possible with most metals. Sharper radii can be formed by a special process of grooving or beading the metal. For simple bending with the neutral axis at 0.5t, the equation for outer fiber strain in percent is:
%e= t 2r+t x 100

Thus, the outer fiber stretch is 33 percent for a bend radius equal to metal thickness. For prepainted products, this amount of stretch could be significant, as any paint film on the surface of the substrate metal must undergo an elongation greater than 33 percent without degradation or reduced performance capability. This may require, for critical applications, roll forming with the surface temperature of the paint above the Tg (glass transition temperature) to allow the surface to flow. However, some metal coated products may not have the option of changing the formability of the metal coating by temperature. In this case, the limit of the bend radii is not set by the formability of the substrate metal but by the tensile stretching limit of the coating on the outer fiber of the bend. The converse compressive deformation without shearing from the substrate is required on the inner fiber of the bend. BLIND CORNERS - A blind corner is a bend or an area of a bend that can not be controlled by direct contact of the rolls. Accurate control of section dimensions is difficult in blind corners. LEG LENGTH - Minimum practical leg length (flat beyond last radius of bend) is three times material thickness. Shorter leg lengths cause nipping of the edge of the material and result in a wave along the edge of the part. SECTION WIDTH - Sections with wide, flat areas in the final part are prone to lose their flatness. Longitudinal ribs are required to mask or hide coil waviness and to remove tendency to oil can (trapped excessive metal which is elastically bowed and can be snapped between deflection above and below the zero position). Roll forming can not remove coil imperfections (loose metal, waves, buckles, full center, etc.) in wide flat areas; it can only attempt to suppress them in the cross section with ribs. NOTCHES AND PUNCHED HOLES - Prepunched holes should be located at least three to five times material thickness beyond where the bend radii will be located. Holes closer to the edge will be distorted during the bending operation. PART LENGTH - The length of the part is created either by precut or post-cut methods. For precut parts, the minimum length should be twice the horizontal distance between roll forming stations. For post-cutoff the final cross sectional shape must be designed to facilitate cutoff.

The design considerations listed above indicate the type of design guidelines available in the lieterature (S-29, G-43, A-14, R-14, J-15, F-3, H-38, G-27, G-28, G-29, G-42, H-2, H-3).

7.2.4 ROLL FORM TOOLING


The greatest amount of art in contour roll forming deals with determination of the required number of roll stations and the design of each pair of contoured rolls (S-29, G-43, A-14, R-14, J-15, F-3, H-38, G-27, G28). The initial design of roll tooling is the development of the flower (Figure 7.2-3). This is a station-bystation overlay of progressive section contours starting with the flat strip and ending with the final desired section profile. The flower also can be obtained by starting with the final profile and unfolding it into the flat strip. Each intermediate profile is a pair of roll forming tools. Two prime considerations in designing the flower are (I) smooth flow of material from the first to the last roll stand and (ii) maximum control over the fixed dimensions while roll forming. The first-smooth flow is easy to visualize. Take a long strip of paper which is the length of the roll forming machine. Hold one end (incoming) flat. Contour the other end (exit) around a pattern of the final form. The gradual change in sweep and curvature would be the ideal progression of the metal flow. If cross sections were taken at equally spaced intervals between the beginning and the end, a flower would be generated. This flower would represent the ideal roll forming with respect to material flow, but would be lacking in dimensional accuracy. Such a progression, however, is continuous and not in discrete steps created by direct contact with each set of contoured rolls. The number of roll sets is limited by cost considerations. Likewise, bending should not take place between roll stands. Any free space (or air) bending can not be controlled as to the location or the dimensions of the resulting deformation. The other extreme would be to form each corner to its completed angle with total or maximum roll contact before leaving the roll set. This would eliminate air forming and blind corners. However, the flow would be a jerky, step-by-step, motion rather than a flowing motion. This stop-go-stop motion would generate excessive stresses which would cause forming problems, as well as require an excessive number of roll forming stands. The compromise between perfectly smooth flow and the controlled deformation is the art required by the roll former designer. The designer starts the forming in the center of the strip and works towards the edge of the strip and impose a tensile stress in addition to the bending stress; this would severely limit forming capacity and greatly increase breakage. The proper sequence is illustrated in Figure 7.2-4a. Some roll formed panels, however, have additional ribs inserted into the panel after the initial forming sequence, as shown in Figure 7.2-4b. These ribs may provide additional stiffening or a cosmetic effect. Now, however, the metal on both sides of the new roll formed ribs is completely locked; metal can not flow into the bend from the free edges. The deformation mode now becomes more complex with a combination of bending and stretch forming. this non-ideal roll forming procedure can be very severe and may exceed the formability of either the substrate metal or its coating or even both. This deformation mode and its corresponding limit are better duplicated in the laboratory by a stretch-bend test than by the more traditional free-bend test (H-43). The designer also attempts to impose less forming in the initial and final roll sets (S-29). Forming in the early stations is limited because material inertia must be overcome. Forming should be slowed down at the later stations so that the section loses its tendency to continue forming at its predetermined rate. This slowing down also helps to eliminate flare on the ends of the section. The design of the final rolls is critical because here overbending is introduced in order to compensate for springback. Methods to develop the flower are suggested in the literature (S-29, G-43, A-14, G-29). Recently, the design of the flower has been assisted by computer computation (R-14, J-15, Y-13).

7.2.5 NUMBER OF ROLL STATIONS


The number of stations is affected by the work metal thickness, composition, and hardness (A-14). Other factors are the complexity of the shape, tolerances, production quantity, available equipment, and special requirements. - For a given shape, the number of work stations increases as the thickness of the work metal increase; the increase is not proportional, however. - Composition and hardness of the work metal strongly influence the number of stations required; an increase in the yield strength increases the number of stations. For this reason, annealed work metals are preferred because the yield strength is the lowest. However, Lders bands (stretcher strains) are undesirable. In some roll forming mills, a roller leveler unit is installed ahead of the first set of rolls to eliminate any yield point elongation (A-14). - Use of an insufficient number of roll stations leads to shape and straightening problems and excessive straightening would be required. - Greater production speed often is obtained by the addition of more roller stations. The instantaneous strain rate at each roller station is therefore held relatively constant because less deformation is being generated at each station.

7.2.6 FORMING SPEED


Speeds in contour roll forming can range in the extreme from 1.5 to 800 ft/min. (0.5 to 244 m/min). The most common speed range is 80 to 100 ft/min. (24 to 30 m/min). The following factors can influence the optimum speed (A-14): 1. Composition of the work metal 2. Yield strength or hardness of the work metal 3. Thickness of the work metal 4. Severity of the forming operation 5. Cutting finished shapes to length (speed of shear) 6. Number of roll stations (strain rate in each) 7. Required auxiliary operations (time for each operation) 8. Use of a lubricant (coolant) An example of the low speed end of the range would be forming of titanium into a complex shape. The high end of the range would be production forming of aluminum or low strength steel in a thickness less than 0.035 (0.9mm) inch with mild forming severity and long lengths greater than 80 feet (24m). Such high speeds would require additional sets of driven rolls and less incremental forming per roll station.

7.2.7 INFLUENCE OF WORK METAL COMPOSITION AND CONDITION


The effect of work metal properties and formability on contour roll forming is generally the same as other modes of forming (A-14): - For a constant thickness, different work metal composition, initial yield strength, and rate of strain hardening will require different power capacities, number of stations, roll material, lubrication, and speed. - When work metal thickness and tooling remain unchanged (common condition), substitution of a more formable metal is seldom a problem with respect to formability limitations. In general, a more formable metal can withstand the shape changes at each roll stand. An example would be forming of aluminum in

tools designed for low carbon steel. The finish station rolls may have to be reground, however, to avoid overbending of the section so that final dimensional tolerances can be met. - Substituting a metal of higher strength or less formability, however, most likely would require changes in the roll pass sequencing; certain section changes could not be tolerated by the higher strength, less formable metal. - Another common change for the roll forming industry is from carbon steel to stainless steel. Problems are often encountered in adequate power and machine strength. While the metal itself may have sufficient formability, the roll former may not have sufficient power to impart the required deformation and maintain production speeds. In addition, insufficient machine strength (rigidity) can cause elastic deformation of the roll stands and housings and therefore cause changes in part dimensions when higher strength metal is formed. For example, for identical shape and thickness, type 302 stainless requires twice the power needed to form 1010 carbon steel. - Higher strength materials require an additional overforming to compensate for springback. - Highly cold worked metals often have residual stresses which cause straightening problems, especially when forming unsymmetrical shapes. This is corrected by adding more work stations to decrease the amount of forming in each station.

7.2.8 LUBRICATION
The primary purposes of a lubricant in contour roll forming are to insure good tool life and protect the metal/coated surface. A recent paper lists twelve important areas to consider to eliminate lubrication mistakes, as well as other problems, in roll forming (I-8): 1. Accelerated tool wear 2. Metal pick up on rolls 3. Surface conditions of purchased material 4. Compatibility of lubricant vs. material 5. Choice of application technique 6. Understanding the secondary operations 7. Unusual lubricant properties 8. Proper startup and operating procedures 9. Preparation of the lubricant 10. Maintenance and control lubricant 11. Secondary metal forming lubricants 12. Maintenance of roll forming equipment All of these topics deal with the protection of the tooling and the work material especially if it is coated with a paint or some other easily damaged surface. For example, emphasis is placed on flushing the surface of the rolls to remove pickup of metal fines or scale from the work metal. If this is not done, scoring and galling will follow. The flushing also is used to reduce the ambient temperature of the rolls. The forming done in roll forming bending appears to be unaffected by lubricants, or the lack of them, in a well designed roll forming operation. In fact, many of the roll forming operations are performed without a lubricant for those parts where surface quality is not critical. In yet other areas, such as the appliance industry, prepainted metal is received with only a wax coating over the paint. No additional lubricant is applied because the plants do not want the extra step of lubricant removal. Here extreme care is taken in the roll forming operation to preserve the quality of the incoming painted surface.

7.2.9 BLANK WIDTH CALCULATIONS

The width of the strip (blank) required to roll form a given shape can be calculated form the geometry of the proposed part (S-29). While this strip width calculation is theoretical, only slight modifications will be required when the part is actually formed. The calculation process begins by dividing a complex part (Figure 7.2-2) into segments each segment being described by a single shape. Each straight line unit is a segment. Each curved element of constant radius is a segment. The example in Figure 7.2-2 has nine segments. The calculation is based on the length of the neutral axis. Therefore, the neutral axis length of a straight segment is its actual length. The neutral axis length calculation for curved sections is called Bend Allowance. This calculation is more complex than calculations for straight segments. METHOD ONE (S-29) Calculation of the Bend Allowance (BA) for a curved section:
BA = R 57.3

BA = Bend Allowance (inches or mm) R = Bend radius (inches or mm) = Angle through which material is bent (degrees) For an inside bend radius greater than two times material thickness: R = Ri + 0.5t where Ri = inside bend radius (inches or mm) t = metal thickness (inches or mm) This formula is equivalent to the neutral axis lying in the center of the strip thickness. For an inside bend radius less than two times material thickness: R = Ri + 0.4t This represents a shift in the neutral axis closer to the inner radius. This reduces the length of the neutral axis but will increase the percent stretch required in the outer fiber.

METHOD TWO (S-29): Instead of only two different positions of the neutral axis, method two varies the position of the neutral axis according to the ratio of the inside bend radius and the material thickness (RA). BA = [(txP) = Ri] (0.) where BA = Bend Allowance (inches or mm) t = Material thickness (inches or mm) P = Material thickness percent Ri = Inside bend radius (inches or mm) = Angle through which material is bent (degrees) 0.01745 = 1/57.3 used in Method One.

For RA (ratio inside bend radius to material thickness) less than one: P = (RA x 0.04) + 0.3 For RA greater than one: P = (RA 1)0.06 + 0.34 with a maximum value of 0.45 for any RA > 2.83. These calculations have the neutral axis varying from 0.3t for a tight radius to 0.45t for a generous radius. These calculations can be easily computerized.

7.2.10 SUMMARY
The forming mode of a well designed contour roll forming operation is bending, both under the contour rolls at each station and between each station. This bending deformation allows steel with low formability in traditional stretching operations to be contoured into complex shapes without breakage. One example is the roll forming of very high strength automotive door beams. The bending mode of deformation also permits calculation of blank width based on constant length of the neutral axis through the bending process. This single forming operation allows rules of thumb to be developed for proper design of roll formed sections, as well as making this forming operation amenable to computer design. Figure 7.2-1 A roll formed part developed in twenty forming stations. Geometry is generated by a

sequential bending operation (S-29).

Figure 7.2-2 Cross-section of a part with the calculated "developed" strip width (S-29).

Figure 7.2-3 Flower development of a lock-seamed tube showing progressive section contours starting with the flat strip and ending with the final section (S-29).

Figure 7.2-4 (a) Proper development of roll forming sequences requires innermost sections to be formed first (numbered 1) so that metal can flow into the forming zone before more outboard sections (numbered 2, 3 and 4) lock the metal. (b) Sometimes stiffening ribs (numbered 5) are added last - these are limited in design and depth because the deformation mode is now stretch-

bend.

The concept of the Forming System has gained acceptance over the last decade. It highlights the fact that any one component of the system can be analyzed only in terms of the remainder of the system. Formability parameters can be specified only in terms of a specific stamping. For example, stretchability can be discussed as being directly related to the work hardening exponent or n value. The absolute value of n required to successfully form a stamping depends on the stamping design plus knowledge of the lubricant, press adjustments, and other die parameters, such as draw bead design, blank size, die radius, etc. This section describes three areas where the Forming System and the interactions among the system components have been considered. These areas are Severity Analysis in the Press Shop, Mathematical Modeling, and Springback.

8.1 Severity Analyses


8.1.1 GRID ANALYSES
Analysis of grids on the surface of a deformed stamping is an important press shop analysis technique. Grids are placed on the blank prior to deformation. After forming, measurement of the deformed grid provides the magnitude and direction of the principal strains over the entire stamping, as well as the flow of metal from one area of the stamping to another. Grids measured on a series of partially formed (incremental stampings) show the relative metal movement within the die as a function of punch travel. In addition to defining the actual strain states in the stamping, the measurements from the grids are plotted on the Forming Limit Diagram (Section 4.2) to establish the severity of the deformation. A typical analysis is shown in Figure 8.1-1. here the strain states, strain magnitudes, strain distribution, and strain severities are shown in a concise and easily interpreted method (F-2). This analysis shows that the entire wall of the corner is critical. Another type of analysis is shown in Figure 8.1-2. Here the severity of the stamping being analyzed increases dramatically as the maximum percent reduction is approached (J-13). The most common grid geometry is the circle grid. The circle patter is used in press shops throughout the world. The details of the circle grid analysis system have been extensively published. Booklets on the subject include A-12, D-18, K-10, M-36, G-13. Other literature includes A-7, C-7, F-2, G-2, G-16, H-14, H-

16, H-17, H-18, I-7, J-13, J-14, K-24, K-26, K-28, K-29, K-30, K-48, S-28, L-6, M-6, S-41. Therefore, the circle grid analysis technique will not be amplified here. Early grid studies utilized square patterns because the grids generally had to be hand scribed; printed grids would be scraped off the surface of the sheet metal during forming. Circle grids were difficult to scribe. Circle grids became popular only with the advent of electrochemical marking or etching of the grid into the surface of the blank from a stencil containing the circular pattern (D-18). Electronic grid measuring systems are now available (H-5). Direct measurements of principal strains from the original square grids were difficult because of the shear component of the deformation required calculation of the principal strains through formulas. Today, analyses of square grids are reappearing. The square grids have certain advantages over circles (T-11). First, the square pattern allows a mesh to be drawn over the sheet surface which will cover the entire surface of the sheet without either gaps or overlaps in the coverage. Second, speed of measurement and flexibility of choice in the reference grid are available, especially with the assistance of computer based, automatic digital coordinate measuring (optical scanners) devices, data acquisition systems, and analysis/computational programs (D-1). Third, the square grid is required to analyze the influence of the strain path when the deformation is non-coaxial. The final measured shape change of the circle does not dictate the deformation mode experienced by that circle during the entire process. The same shape change can be achieved by an infinite number of monotonic deformation paths. Unlike the circle, a square grid shows both the stretch component plus the rotation of the element from its initial direction. Several recent papers have used the square grid along with certain simplification procedures to develop a viable press shop analysis tool (C-10, C-12, N-13, T-12, S-32, S-33). The simplification process is based on the assumed pure homogeneous deformation mode. Studies of springback and other elastic/plastic deformations require the analysis of small strains. An example is the deformation and deflection of an automotive hood panel. Here a large circle ranging from 1 to 10 inches (25 to 250 mm) in diameter is scribed in the panel surface to improve the measurement accuracy of the small strain levels (D-18, K-10). However, the problem of determining accurately the direction of the principal strains from the deformed circle is magnified at the low strain levels. One solution is to use the brittle lacquer technique (where cracks occur normal to the largest principal strain direction) to gain additional information for strain state analysis (U-1). Another analysis technique is the use of Moire patterns to observe sheet metal deformation (Y-11). Finally, the bonding of wire strain gages to the surface of the sheet has been used (A18). The advantage here is that high speed recorders can be used to record the strain history for delayed analysis; such histories are obtained with full production conditions of press speed, temperature, etc. However, the square grid analysis can be used effectively for the low strain measurements of automotive body panels. Excellent work was performed by Nihill and Thorpe (N-13). They developed a commercial software program entitled "SIROSTRAIN". This computer program computes the first and second (major and minor) principal surface strains and their orientations from the measured distortions of square grid patterns. The results were displayed in the form of a series of maps showing contours of equal strain magnitude and the orientation of the first principal strain. These maps for automotive body panels are shown in Figure 8.1-3 for an outer hood panel, Figure 8.1-4 for a right-hand front fender panel, and Figure 8.1-5 for right-hand rear quarter panel. Once the measurements are stored in the computer data base, additional analyses can easily be performed. For example, the strain states for each panel and the strain levels can be displayed on a strain space axis similar to that used by the Forming Limit Diagram (Figure 8.1-6). Note that most strains are positioned around the plane strain axis.

8.1.2 STATISTICAL ANALYSES


The Forming System is amenable to many types of statistical analyses. Standard Statistical Process Control (SPC) techniques can be used to analyze various processes to determine if they are in or out of control and to determine the process capability. Examples could be the thickness of the incoming steel, the dimension of the blank, the thickness of the applied lubricant, and other single measurements. With

the installation of proper sensors/monitors, other SPC analyses can be done. Key sensors currently being installed in the press shop are load monitors on all four corners of both the inner and outer rams. These are useful for initial settings after die changes and to assess the consistency of the settings over an extended press run. This, of course, helps control metal flow in a uniform manner. On the other extreme, the Taguchi method is used to design meaningful experiments which can be run in the press shop with a minimum of disruption to production. Here the sensitivity of the Forming System to specific variables can be evaluated at any given point in the history of a stamping. A need exists, however, for an analysis systems which will be sensitive to the interaction of all variables in the Forming System, can be sued to quantitatively define the severity of the system (without any breakage), will monitor the Forming System over the life of the tooling, can be used to determine the cause of system variability, and can evaluate the sensitivity of the system to any variable. One promising system for achieving these goals is called Statistical Deformation Control or SDC (K-11). With SDC the status of the Forming System is obtained through ultrasonic measurements of stamping thickness in the critical zone (K-11). Converted to a thickness ratio, the severity of the stamping is obtained from a thickness Forming Limit Diagram. Standard SPC techniques are used to obtain a control chart and to analyze the system with respect to the control chart. Breakage of the stamping is neither required nor desired. Various studies can be conducted to determine the cause of system variability and sensitivity.

8.1.3 SHAPE ANALYSIS


Chrysler Engineering has pioneered an analysis system which relates formability demands by the part configuration to the formability capacity of the various metals (K-4, K-5, K-6, K-7, V-2). The formability capacity of metal is first evaluated in the laboratory for conditions of pure stretch and pure draw. Using these two measurements as end points, a forming line is constructed (Figure 8.1-7) to represent the maximum amount of forming allowed (Forming Ratio) for any combination of stretch and draw. Each candidate material, or different samples of the same material, can be so characterized. The second step is to measure certain parameters (such as blank size, punch diameter, stretch-draw boundary, etc.) on the actual formed stamping. These parameters are used to calculate a percent draw and a forming ratio of the panel under study. Panel severity is a comparison of the actual forming ratio assigned to the stamping and the allowable forming ratio of the material (See Figure 8.1-7). The distribution of strain influences the stretch-draw boundary and therefore becomes part of the severity calculations. Although qualitative results can be obtained from blueprints, quantitative measurements require a formed stamping with circular grids. A cooperative program of the American Deep Drawing Research Group (A-7) evaluated a number of parts ranging from stretch to draw. The shape analysis severity rating and the FLD severity rating showed a linear relationship. The shape of the stamping often determines the relative amounts of stretch and draw required to form the stamping. Any analysis of shape which proportions the stamping into stretch and/or draw can be used to estimate the dependence of the stamping to the formability parameters n and r (Figure 8.1-8). Such a diagram can be constructed for any given stamping if a series of steels with varying n and r values is available.

8.1.4 CASE HISTORIES


Much of the actual sheet metal forming data available today is in the form of case histories. A variety of data presentation techniques are found in the literature. These range from reports (Figure 8.1-9) on specific part trials with different steels (A-8) to a strain composition diagram (Figure 8.1-10) of the whole surface of a car body (I-6, K-16). Strain frequency histograms are becoming common in the literature (Figure 8.1-11). Likewise, the various strain states and strain levels are being coded (Figure 8.1-12) and

then used to provide a visual presentation of the deformation of panels. One example (Figure 8.1-13) shows the development of the strain for various stages of the punch stroke for two different punch configurations. Obviously such data are invaluable to the press shop. The data are both visual and quantitative. They provide a record of the deformation in the stamping for each trial variable. Trends can be observed. More important, permanent records can be kept for future reference. These case history records also serve an important purpose as experimental verification for mathematical modeling calculations (Section 8.2).

8.1.5 SUMMARY
Severity analyses currently utilized by the press shops depend upon informaiton gained from the stamping itself. Circle grids allow the major and minor strains from critical locations within the stamping to be characterized in terms of forming severity. Statistical analyses provide a systematic method of tracking the severity of a stamping throughout its production life cycle. Shape analysis relates the relative amounts of stretch and draw within a stamping to their respective limits determined from laboratory tests. In each case, the analyses are postmortem in nature. The formed stamping reflects the summation of the Forming System components. Only then after actual forming of the stamping can an estimation of the forming severity be made. The Forming System is so complex that results of changes in the components currently can not be predicted. Until improved mathematical modeling techniques are created, the existing severity analysis techniques can be utilized to reduce the severity of stampings. Figure 8.1-1 Strain analysis shows the entire wall of the pan was in the critical zone with high probability of failure (F-2).

Figure 8.1-2 The severity rating increases dramatically as the maximum percent reduction is

approached (J-13).

Figure 8.1-3 Strain map for an automotive outer hood panel showing contours of equal strain

magnitude and orientation of the maximum strain (N-13).

Figure 8.1-4 Strain map for an automotive front fender panel showing contours of equal strain

magnitude and orientation of the maximum strain (N-13).

Figure 8.1-5 Strain map for an automotive rear quarter panel showing contours of equal strain magnitude and orientation of the maximum strain (N-13).

Figure 8.1-6 Strain distributions shown for various automotive panels as a function of major and

minor strain ratios. The numbers for each ring display the occurence frequency (N-13).

Figure 8.1-7 The shape analysis technique to determine an allowable forming ratio based on the

percent of draw in the stamping (K-6).

Figure 8.1-8 Different parts are sensitive to different formability parameters depending on the ratio of stretching to drawing (B-10).

Figure 8.1-9 Strain analysis report commonly used in press shops around the world (A-8).

Figure 8.1-10 Strain composition diagram for the whole surface of a car body (I-6).

Figure 8.1-11 Classification of forming strains into categories based on major and minor strain combinations (I-6).

Figure 8.1-12 Frequency diagrams for the major and minor strains found on different panels in a typical automobile (I-6).

Figure 8.1-13 Changes in the punch contact zones created when stampings are made with

different die shapes (I-6).

8.2 Mathematical Modeling


8.2.1 HISTORICAL PERSPECTIVE
Mathematical modeling for sheet metal formability is thought to be a recently developed science, but, in fact, a paper by Sachs provided a mathematical model for the deep drawing a cylindrical cup in 1935 (S2). Since that date, :improvements in stamping have been due not so much to any enhancement of die design skills but to a continued improvement in material uniformity and to recent discoveries in tool materials and tool surface treatment and the potential for reducing die costs due to electro-discharge machining and numerically controlled contour milling machines..It can be expected that developments in modeling will lead to better understanding of sheet forming and, as a consequence, better tools and better forming machines, but as yet this has not been fully recognized (D-25).

The axisymmetric or cylindrical deep drawing process has occupied a major role in the early development of mathematical modeling. The deep drawn cup contained the major sheet metal forming modes drawing, stretching, and plane strain. In addition, the geometry was sufficiently simple for analytical purposes especially in the precomputer era. Many early analyses were made by Swift (S-45, S-44). Woo followed with numerical modeling using finite difference methods for punch stretching and drawing (W-27, W-28). The initial modeling studies in sheet metal formability were conducted as a means to understand the behavior of the metal as it was transformed from a flat sheet into its final configuration. Much of the early work was aimed at providing a mathematical model of a simple sheet deformation process that could be verified in the laboratory. For a given geometry of tooling, initial blank shape, and a constitutive relation for the material, the objective was to obtain mathematically a result that agreed reasonable well with experiment. The goal of current mathematical modeling activity is different from the early work in the field. Instead of trying to duplicate mathematically an existing stamping, math modelers now are trying to determine for a given final stamping shape how this stamping ought to be produced in the press shop and then to attempt to devise a process and tooling to achieve the shape. The ultimate goal would be to eliminate any soft tryout tooling and to design directly the final tooling which would produce quality stampings with no modifications required during die tryout. In this concept, all tryout, modification, adjustment, and optimization of the forming process would be conducted within a computer program and the final design then committed directly to another computer responsible for the CAM (Computer Aided Manufacturing) of the final die set. The driving forces for mathematical modeling are many and include: - Development of the necessary forming configurations (thickness and final flow stress) to comply with inservice part requirements. - Selection of correct Forming System variables (die design, material parameters, lubrication, press characteristics) and evaluation of system severity before the design is committed to tool production. - Maximum utilization of the developed part surface which already has been mathematically described in the computer data base during the previous styling processes. - Reduction in time and cost from concept to first production for sheet metal stampings.

8.2.2 ANALYSIS CONCEPTS


The central goal for modeling of production stampings is the generation of the dies necessary to form the stampings. This involves translating the geometry of the part as defined by the stylist into the final sheet metal stamping. This deformation of the sheet is accomplished in the die. Because sheet metal forming is so complex, the dies used for this deformation may bear little resemblance to the final part geometry. Preforms, binder areas, restrikes, flange and trim, springback corrections, and even reverse forming (where the initial stamping is turned inside) are intermediate steps in the contortion of sheet metal into its final shape. In the extreme, 75 percent of the sheet metal used in forming a stamping may be trimmed and thrown away as offal or engineered scrap; an average for all automotive stamping is around 35 percent. A great deal of time and man-hours currently are required for the designing, manufacturing, and modifying of the dies. This process is necessary for the production of quality stampings. Unfortunately, techniques for this designing have been inadequate. The personal experience of the skilled workers and the sum total of past trial-and-error development of similar stampings have been the sole base to accomplish this very complicated task (K-23).

The computer has the potential to replace this trial and error process with a formalized, mathematical analysis of great complexity and speed. Each section of the stamping can be modeled. The interactions between sections can be formalized as one section becomes the boundary conditions of the adjoining sections. All parameters of the Forming System can be included. Changes can be implemented in minutes instead of the days required for required modifications to the hard tooling. Most important, numerous what-if scenarios can be research to optimize the design before committing to the final design parameters. Generally, three levels of computer modeling have been identified each level using models of increasing complexity and accuracy but decreasing interactivity (B-9). These levels are: - The first level is based on geometrical concepts similar to the one presented by Toyota (T-1). The results given by this code include the initial punch/sheet impact position, the deep drawing depths of the part, and an estimate of the elongations in the formed part during the process. The first stage of the analysis is the numerical determination of the shape flange over the die cavity corresponding to the binder wrap stage. The formulation is based on the linear elastic membrane theory of thin shells; the boundary conditions are imposed displacements. Essentially, this method closely corresponds to the empirical approach used by stamping designers (B-9). This method is not convenient for accurate stress/strain evaluations and only estimates the elongations in the part. - The second level uses simulation codes which are a faster response method. The strain distribution is obtained by using a direct iterative systems for calculations between the known final geometry of the part and the initial flat geometry of the sheet. This technique was previously developed by Lee (L-5, C-13) for two-dimensional and axisymmetric configurations. The computational procedure uses a simplified variational method. This is derived from the principle of minimum potential energy to determine the initial nodal positions in the mesh corresponding to the flat configuration. - The third level has an analysis scheme for the complete calculation of stress and strain distributions at any stage of the deep drawing process. This uses large elastic-plastic Finite Element Modeling of sheet metal forming with contact boundary conditions. The basic hypotheses to establish the formulation are membrane thin shell theory, plane stress condition, isotropic elasto-plasticity with Von Misess hardening criteria, and contact boundary conditions with Trescas friction law (B-9). This third level modeling is the most difficult to develop, both on the theoretical and numerical aspects, and needs to be improved in the methods used to incorporate friction boundary conditions and the mechanical behavior of the sheet metal used. An ideal designing concept using CAD (Computer Aided Design) is shown below and would replace the current trial and error process for creating the required die design (T-2, H-27): A) Data base with part shape B) Part shape analysis C) Tipping position D) Step draw E) Blankholder surface F) Draw bead G) Evaluation of press forming severity H) Blank shape I) Stamping die data base J) Final drawings or CAM system input. Each step is described in greater detail below.

8.2.3 ANALYSIS STEPS

A) Data Base With Part Shape The starting point for any analysis is the mathematical description of the required part or the specification of the required surface. The normal sequence for styling has been development of the clay model, the digitizing of the clay model, mathematical smoothing of the design lines into more consistent equations, and then correcting the clay model. Ideally, however, the styling should be done using a computer. From this computer design the clay model can be automatically machined for visual checking. This sequence eliminates the extensive digitizing procedure and correction of the clay model because the mathematical description of the surface was created as the styling progressed. Sheet metal parts can be divided into two categories. The first are those surfaces which are developable and singly-curved. Developable surfaces can be developed by combining formalized, three-dimensional mathematical shapes, such as cones and ellipsoid. These parts can be formed by roll forming, press brake bending, folding, and flanging. These shapes can be described by straight line generators and are single-parameter surfaces (D-25). The descriptions of such surfaces do not present great problems. The second category of surfaces are those formed by stretching and drawing so that they become nondevelopable and doubly-curved. These stampings usually are irregular and highly complicated in shape. Automotive outer body panels are in this category. Even though these panels have smooth contours, the actual shape can not be fitted by formalized, three-dimensional, mathematical shapes such as cones, toroids, and ellipsoids. Automotive inner body panels span both categories. The panel can be broken down into a number of simple developable shapes, but the interconnection of these elements is complex and mathematically difficult to describe. A mathematical description of the part surfaces is one of the important issues in mathematical modeling and will continue to be the target of detailed future research. An important issue in surface description, however, must be addressed now that of standardization. Unlike flexible human beings which can interpret and digest various systems of presentation, computers are very rigid in the format in which input data must be submitted. Even more important, workable CAD/CAM systems will extend rapidly to contract stampers world wide. Standardization of data from a variety of requesting companies. A proliferation of system formats will either restrict the ability of various companies to bid on available jobs or force them into the expensive option of being equipped with numerous systems. The need for a standard language for mathematical surface description is great. B) Part Shape Analysis This step represents the heart of the mathematical modeling process. The mathematically described surface is entered into the analysis program and is modified to provide the necessary output. The output of the analysis can vary from one program to another. However, a typical output might include: a) a mathematical mesh of the deformed grid (Figure 8.2-1), B) strain history plots (Figure 8.2-2), strain level plots on Forming Limit Diagrams (Figure 8.2-3), frequency histograms of strain ratios (Figure 8.2-4), What-if-scenarios (Figures 8.2-5, 8.1-6), or pictorial displays of forming severity (Figure 8.2-7). The two most common analysis procedures are finite element analysis (FEA) and geometric modeling. Finite element analysis (FEA) methods are the most popular approach to solving sheet metal forming modeling problems (W-4). These methods describe, by a sequence of discrete steps, the manner in which a component is formed. The calculations are continued until a final shape is realized or until some failure condition (wrinkling, tearing, localized necking, unacceptable thinning, etc.) is met. If the model shows that the part can not be formed, the initial conditions are changed and the calculations are repeated until a satisfactory part is achieved. The initial conditions can include all the material formability parameters (strength, work hardening, anisotropy, temperature, strain rate effects, etc.), interface friction, punch/die geometry, and other

variables in the Forming System. To date, some success has been achieved in analyzing parts of simple geometry, particularly for axisymmetrical components (D-21). The FEA method is less satisfactory when more complicated shapes are considered (C-11). Extensive computation time is required, even for parts of simple geometry. At the present time the technique is not suited for interactive computer-aided design. The time level may be reduced to a practical level for the new supercomputers. A simpler, more approximate approach to mathematical modeling is geometrical modeling. In this process, a sheet metal component is transformed from a flat sheet into a nondevelopable surface without a change in thickness. In other words, each element would be deformed by plastic shearing so that its area remains constant. Although few practical forming processes occur in this way, many traditional die design rules are based on a similar assumption, namely that the overall blank area remains unchanged during a stamping operation (C-11). This assumption provides a useful basis for tool and blank design. Thus, geometrical modeling achieves the transformation of a flat sheet to a developable shape by bending only (constant area at the neutral axis). This means that the strain path is one in which the major and minor strains are equal and opposite. With this assumption, the transformation from a flat sheet becomes a kinematic problem rather than one of continuum mechanics and it is intrinsically easier to solve. Although this mapping method is approximate, it does not facilitate the analysis of complicated surfaces (D-24). This technique was used to predict the deformation of a square grid during forming of an automotive rear deck lid inner panel (D-25). The predicted deformation agreed favorably with the measured deformation. More important, the computation time was less than ten seconds on a microcomputer; this makes the geometrical modeling approach useful for an interactive computer-aided design system available to many more die design shops. A third method is the shell theory (D-25). For axisymmetric shapes, the surface can be divided into a number of shell elements of spherical, cylindrical, or toroidal shape. The advantage is that this system can be used in a number of different ways. For example wither force or thickness can be used as initial boudary conditions and iterated to obtain a reasonable solution. Wood et al have compared analysis of thin sheet forming using the viscous shell formation and a solid mechanics based method (W-29). This paper concluded that the viscous shell approach is computationally faster than the elasto-plastic counterpart, but at the expense of the exclusion of elastic behavior. The elastic behavior becomes important, however, when the calculations must correct for springback (D-5, M-24). Whatever the computational procedure, the solutions are only as good as the description of the material behavior. A variety of material phenomena have been found which require modification of standard plasticity formulations for accurate representation of the material as it affects the formability of materials. In the paper by Wagoner the modifications included non-isotropic hardening, complicated strain-rate sensitivity formulations, transient behavior along nonproportional loading paths, and temperaturedependent flow (W-1). A particularly difficult description of material behavior is that of coefficient of friction needed in most of the analyses. Stine (S-38) reported that agreement between the computer model and grid analysis results was substantially improved when experimentally determined friction coefficients were used as input to the computer model. In contrast, the use of arbitrary estimated coefficients can have a very detrimental effect on the ability of the computer model to predict actual results. The coefficient of friction depends not only on the material and lubricant being used but also the deformation mode and tool material (See Section 5 Lubrication). C) Die Tipping Position Automotive stampings are formed with the sheet metal blank approximately in the horizontal position. The specific orientation of the final stamping geometry relative to the plane(s) of the blankholder is a critical decision in the die design process. Two extremes are shown on Figure 8.2-8. Depending on the angle of

the die, the deformation can be made (a) by a plane strain stretch without required metal movement from the blankholder or (b) by a bend-and-straighten operation with metal movement from the blankholder. The plane strain stretch deforms the metal in free space and avoids dragging the metal (with potential scratching) over the die radius. However, the strain is highly localized over the punch and balancing of the deformation forces is required to avoid sliding of the sharp punch across the ace of the stamping. Consideration must even be given to the dimensions of the stamping to allow the stamping to be removed within the confines of the die opening. Obviously, computer modeling of the forming process is the ideal method for studying the tipping angle of the stamping. To evaluate this with real tools requires major and time consuming modifications ot massive sets of tooling. D) Step Draw The draw wall is the location of many failures in sheet metal stampings (S-21). Failure in the wall has been divided into three zones (Figure 8.2-9) punch shoulder, center wall, and pass-through-drawbead. Metal in the straight sides of the stamping generally undergoes plane strain deformation, which allows for easier modeling. The forces imposed on this area are created by the blank geometry, the blankholder configuration, the draw beads, die radii, and interface friction. Therefore, the step draw is evaluated in conjunction with the next two analysis stages: E) Blankholder surface and F) Draw Bead (H-27). Successful forming of the stamping wall depends on the tensile strength of the steel being stretched in a plane strain deformation mode (S-21). This factor determines the capacity of the wall to withstand necking while generating sufficient force to pull the blank length through the blankholder/draw bead restraining system. Shiokawa states that it is possible to estimate the amount of tension, elongation, and movement of material in each part in the process of panel forming, and to predict the occurrence of fracture. (S-21). E) Blankholder Surface Another difficult decision phase in die design is the development of the blankholder surface (T-4). Blankholder surfaces seldom are designed in a single, horizontal plane. Instead, the closing of the binder ring causes a deformation of the blank into a preform configuration (Figure 8.2-10). The preform has two goals. One is to shape the metal to the general configuration of the punch (wrap) so that the punch does not travel an extended amount while riding only on the sharp edge or character line of the punch. Instead, the general contour of the punch is preformed into the blank so that contact over a major portion of the punch is accomplished with little additional travel. This can be seen in Figure 8.2-10. The punch travels well below both edges of the blank (and especially the right hand edge) before contact is made; then contact over the entire surface is rapidly accomplished. The other goal is to avoid locked metal between two initial contact points on the punch. This locked metal will be the site for potential buckles, low spots, and other surface contour defects. This latter determination is difficult to predict without mathematical modeling. When the blank is preformed by the binder geometry, various sections of the blank undergo different amounts of elastic and plastic deformation. Combined with residual stress patterns, the contours of the sheet surface within the binder opening can deviate greatly from the geometries of the binder surfaces. This changes the pattern of contact sequences between blank and punch. Trapped metal can occur in totally unexpected areas. A number of papers have been devoted to the mathematical modeling of this binder wrap (t-2, M-28, I-6). F) Draw Bead Laboratory tests utilizing draw beads are discussed elsewhere (Section 4.3.4.6). However, the deformation mode occurring within the draw bead itself has been the subject of mathematical modeling (N-14, S-21, L-7, L-9, W-5, Y-4). The experimental data and the mathematical modeling have shown the importance of the draw bead system both as a design option and as a process variable. When used as a design option, difficulty has been encountered building the exact bead geometry into the die and then maintaining the bead geometry over the life of the die. After changing blankholder force, the

next die modification used to make the steel function properly is a change in bead geometry weld up, grind down, or both. The draw bead then becomes an unknown and uncontrolled variable. Future stamping design incorporating complete blank locking to achieve pure stretch forming will eliminate the draw bead as a process variable (Section 7.1). G) Evaluation of Press Forming Severity Forming severity can be interpreted in two ways. The most common relates the calculated strains to the Forming Limit Diagram. Details of the Forming Limit Diagram as a determinator of forming severity are provided in Section 4.2 (Forming Limit Diagrams) and Section 8.1.1 (Circle Grid Severity Analyses). If the strain level exceeds the Forming Limit Diagram, localized necking and failure are probable. For strain levels just below the forming limit on the Forming Limit Diagram, the stamping has an insufficient safety factor. Most mathematical models use the Forming Limit Diagram as the terminal point for deformation calculations (D-25, S-38, B-5). However, other parameters are used, such as the Mean Section Length Ratio and Local Elongation Ratios (H-27). Even though the Forming Limit Diagram began as a quantitative description of the onset of localized necking, the FLD also has become a convenient method of portraying all possible strain measurements in sheet metal forming. As shown in Figure 8.2-11, the FLD now is used to depict the allowable strain space and becomes a graphical method of presenting both experimental and computer modeled results (S-38). Another definition of forming severity is the ability of a die set to produce a stamping to the exact mathematically described surface without low spots, wrinkles, buckles, waves, loose metal, etc. Thus, stampings can be ranked in severity according to the ability to be produced without these cosmetic defects. On this basis, quarter panels (low spots above the wheel house opening), decks and lids (large flat areas which are difficult to pull tight), and doors (buckles emanating from the door handle pocket) are examples of severe forming. H) Blank Shape Superimposing the trim line on the final stamping shape now determines the initial blank shape. Opportunities exist here to optimize the blank shape, such as reducing the blank dimensions, changing a contoured blank into a rectangular sheared blank, improved nesting, and other cost saving factors. Mathematical modeling allows the user to verify that the blank shape changes will not prevent the final stamping requirements from being accomplished. I) Stamping Die Data Base The final stage in the modeling analysis is the storage of the complete mathematical description of the finished stamping, the die geometry, and all other components of the forming system. The structure of this data base system must be designed to ensure easy management of the mass data, rapid recovery from system failures, and quick response to command execution (H-27). A whole field of expertise is being developed in this area of massive data generation, storage, retrieval, and flow between different computers. Discussion of this field is beyond the scope of this report. J) Final Drawings Or CAM System Inut Completion of the mathematical analysis will provide (when the system reaches its full potential) a complete description of the tooling surfaces (which all be different from the stamping surface because of springback, subsequent forming operations, etc.), the binder surfaces, draw beads, etc. This information can then be translated into traditional drawings or used as an input for the Computer Aided Manufacturing. This CAM can be an in-house activity or the data can be electronically transmitted to an outside contractor for the actual construction of the tooling. Computer Aided Manufacturing and other programs for manufacturing the tools for producing the stamping are beyond the intended scope of this report.

8.2.4 CASE HISTORIES


Mathematical modeling ha the ultimate goal of designing the final stamping configuration and the corresponding die geometry such that the stamping produced will have strain states and strain levels within well defined boundaries. To do this the modeling system must be able to calculate the strain values at all stamping locations. Thus, an important test of each proposed modeling system is to compare the calculated strain profiles with experimental strain values measured on formed stampings. A number of programs have reached that critical stage although the stampings may be simple, axisymmetrical designs. The following is a partial list of the stampings which have been modeled: Stamping 1984 Camaro Lift Window Outer Oven Door Liner Rectangular Punches Hemming Simplified Deck Lid Rectangular Pan Conical Cup Rear Quarter Panel Press Brake Bending Superplastic Rect. Box Roll Forming Channels Channels Deep Draw with Square Blank Wall Breakage Front Fender Hemispherical Zone LDH Dome Non-axisymmetric Cup U-Bend Circular Discs Dryer Inner Door Hemispherical Dome Swift Cup Test
Stretch Flanging

Author Arlinghaus

Company GM

Reference

A-19 Stine GE S-38 Stoughton GM S-42 Makinouchi Riken M-5 Tang Ford T-4 Doege Hannover U. D-19 Logan U ofo Mich. D-19 Shiokawa Nissan S-21 Nagpal Battelle N-1 Ghosh Rockwell G-11 Yuen Lysaght Y-13 Wang Ford W-9 Chakhari Grenoble C-2 Kaftanoglu Mid-east K-1 Hayashi Sumitomo H-10 Yamasaki Mazda Y-2 Ziemloewocz U. of Wales Z-2 Gerdeen Mich Tech G-7 Iseki Tokyo Inst. I-5 Makinouchi Riken M-4 Nakamachi Yatsushiro N-4 Chung RPI C-13 Kim Soul Univ. K-36 Budiansky Berkeley B-21, K-44
Wang GM W-8

8.2.5 THE FUTURE


The introduction of large scale computer graphics systems opens up new possibilities in describing surfaces and their deformation (D-25). The detailed analyses with equilibrium and constitutive equations are the most precise of the numerical models. However, the purely geometric methods appear to be the most useful at the present time (D-25).

The development schedule for introducing computer modeling at Volvo Car Corporation appears to be a realistic summary of the field (K-50). During the last five to ten years styling, pre-production engineering, tool design, NC machining, and dimensional measurement and verification have been applied to Volvo. The future schedule is as follows: - Improvement in modeling techniques for forming processes in general. - Obtaining methods of finding intermediate shapes in multi-stage forming processes. - Determination of metal flow during stamping. - Calculations of the relationship between strain and draw depth. - System for finding the appropriate developable surface for wrapping the sheet around the punch in the blankholder. - Calculation of a force balance to aid in the choice of punch entry angle. - Enhancement of a system for documentation, storage, and retrieval of performance data. In order to optimize the use of modeling methods, our understanding of the relation between process design, product quality, and productivity must be deepened. Computer-aided modeling and analytical techniques also require precise information on mechanical properties and process variables such as friction. Ideally, the process can be reversed, whereby the modeling of the process can be used to develop the material parameters necessary to successfully form the designed part. The resulting constitutive equations can then be related back to microstructural parameters such as texture coefficient, cell size, grain size, etc. (F-13). In addition, these must be readily accessible, hence the requirement for documentation and storage of process data (K-50).

8.2.6 SUMMARY
The goal of mathematical modeling is readily understandable. Instead of constructing tooling to evaluate potential forming problems, the forming process is simulated in the computer. The strain distribution, the forming severity, springback, metal flow, blank shape, binder configuration, and other parameters of the Forming System are analyzed quickly by a computer program. Changes to the Forming System can be analyzed interactively as the search for the best parameter combination progresses. Cost of development tooling and die tryout are eliminated or at least reduced. More important, the lead time from concept to final tooling is radically reduced. The benefits are unquestioned. The key to mathematical modeling is the design of the simulation program. Currently several levels of analysis are under development, ranging from simple geometrical line fitting to complex analyses implementing metal working and anisotropy, interface friction, deformation speed, tooling temperature, and other parameters. The more complex the simulation, the more accurate the results. However, the complex simulations currently require tremendous computer speed, memory, and computation time. While significant progress in mathematical modeling has been made during the last twenty five years, major work remains before the programs are readily useable in the press shop. Figure 8.2-1 A square mesh on the initial blank is mathematically deformed to produce the strains

required the final geometry of the stamping (D-25).

Figure 8.2-2 The strain distribution across the stamping and a historical development of the

gradients can be presented after completion of the analysis (W-29).

Figure 8.2-3 The strain ratios of the various analysis elements are presented in "strain space" (D25). These points were obtained from Figure 8.2-1.

Figure 8.2-4 Histogram of the relative frequency of nodal strains for the stamping shown in Figure

8.2-1 (D-25).

Figure 8.2-5 Various parameters can easily be changed to evaluate many "what if" scenarios (B-5). Here the effect of the work hardening exponent is examined on strain distribution in a) equal

biaxial stretching and uniaxiat tension.

Figure 8.2-6 A change in blank size requires a modification of a radius in order to keep the maximum strain level below critical (T-2).

Figure 8.2-7 Coding of the stamping (original in color) according to strain level (A-19).

Figure 8.2-8 Die tip changes the deformation mode from (a) plane strain stretch to (b) bend-andstraighten.

Figure 8.2-9 A number of different failures can occur in the sidewall or "draw step" (S-21).

Figure 8.2-10 Pictorial view of a binder, blank, and die during the binder-wrap stage (T-4).

8.3 Springback
8.3.1 INTRODUCTION
Sheet metal forming involves the transformation of a flat sheet of metal into a useful shape. A common problem in this transformation is the distortion in the shape of the stamping that occurs when the deforming load is removed or when the stamping is removed from the tooling. This dimensional change is called springback and results from changes in strain produced by elastic recovery. Springback is inherent in all sheet metal forming processes. The amount of springback depends on the interaction of part geometry, friction, material properties, and die design. That is, springback results from the interaction of all components of the Forming System. The importance of springback depends on the end application, quality requirements, subsequent processing steps, etc. For example, excessive or

variable springback in a channel which is subsequently welded with critical dimensional control is an important processing parameter. on the other hand, springback in the back corner of a floor pan may be relatively unimportant and corrective action may be unnecessary. Two basic approaches are used to correct for springback. The first design of the forming process to eliminate the springback through control of the stress-strain patterns generated in the stamping during the forming operation. The second is not to control the springback but to compensate for the springback in the original design of the tooling for example overbending such that the final part will be dimensionally correct. Each of these approaches requires an understanding of the causes of springback and the factors which affect its magnitude. With this knowledge, trial and error attempts at springback correction become more effective. Ideally, a method is required for quantitatively predicting the magnitude and direction of the springback and/or the residual stress distribution after the forming operation is completed. From this information exact springback corrections can be calculated. More importantly, the sensitivity of the springback to the variability within each component of the Forming System can be calculated. A STATE OF THE ART REVIEW ON SPRINGBACK was completed for the AISI in 1986 by Professor Gerdeen of Michigan Technological University and Professor Duncan of the University of Auckland, New Zealand and formerly of McMaster University (G-6). This section will summarize the key segments of that document.

8.3.2 ORIGIN OF SPRINGBACK


Springback is the result of elastic recovery of the sheet metal after deformation. This can be easily shown from the stress-strain curve obtained from a tensile specimen (Figure 8.3-1). In forming operations, the sheet metal is loaded to a stress exceeding the initial yield stress of the metal (point A). This increment of deformation contains two components the elastic strain which will be recovered after unloading and the plastic strain which will remain as the permanent deformation of the sheet. In Figure 8.3-1 this elastic component is shown as B-C and the plastic component is C-D. The elastic strain often is called the recovery strain and is equal to the ratio of the stress at A to the Youngs modulus for the metal being deformed. Therefore, a 30 ksi (210 MPa) stress would produce a recovery strain of 0.1 percent for steel with a Youngs modulus of 30 million. The same stress would produce a recovery strain of 0.3 percent for aluminum with a modulus of 10 million. Keeping the modulus at 30 million, a 30 ksi (210 MPa) stress for yielding a high strength steel would produce a 0.3 percent recovery strain. Therefore, a recovery strain could be calculated from the mechanical properties of any sheet metal being used to make a part and the flow stress to which the sheet metal is being subjected. Using the above analysis, a 60-inch (1525mm) hood produced from low strength steel would recover about 1/16 of an inch (1.6mm). This change in dimension due to elastic recovery could easily be corrected by making the die 1/16 inch (1.6mm) longer. Unfortunately, real world stampings are not formed like uniaxial tensile specimens and therefore do not produce a uniform linear component of elastic recovery. In most sheet metal stampings, the recovery strain is combined with one of several mechanical multiplying factors which tend to greatly magnify the shape change (G-6). To illustrate this important phenomenon, bending is used. In a typical bending operation, the stress distribution through the sheet thickness while still under the forming moment (load) is shown in Figure 8.3-2. The inside surface of the bend has a compressive stress, while the outside of the bend has a tensile stress each of which must recover in the opposite direction upon unloading. Although the recovery strain is very small, significant shape changes can be generated by the mechanical multiplying effect.

This multiplying effect can be illustrated by a simple experiment (G-6). Two strips of thin cards are joined at the top by a thin piece of tape (Figure 8.3-3). The two cards are held at the bottom; very small movements between the finger and thumb cause large changes in the curvature.

8.3.3 CLASSIFICATION
Spring back can be classified into three major categories and several subcategories: 1. Small radius bends along a straight line a. Pure bending b. Tension plus bending c. Channel bending 2. Small radius bends along curved lines 3. Doubly-curved, shallow, smooth panels For the automotive industry the first two categories describe the forming of beams, channels brackets and other structural members. Body panels fall into the third category. The literature describing each of these categories is reviewed in the next section.

8.3.4 SRING BACK LITERATURE 8.3.4.1 Small Radius Bend Along A Straight Line
a. PURE BENDING This deformation mode describes parts formed by roll forming, V-block forming, and folding. The springback is dependent on the yield stress, elastic modulus, and bend ratio. The basic equation describing springback in simple bending is due to Gardner (G-1):

This equation shows that the springback ratio,

is proportional to the ratio of plastic stress to elastic modulus and the bend ratio, R/t. These parameters are shown in Figure 8.3-4. Another description of springback is the shape retention factor give in equation [1]. Here Rs is the radius after springback. This equation of Gardner is applicable only for bending of a beam. Crandall, in his discussion of Gardners paper, (C-16), stated that elastic springback of a curved sheet should be treated as a plate instead of a beam and therefore the plate stiffness should be used instead of the beam thickness. In addition, bending of a wide plate will involve a term for the elastic Poisson ratio.

This equation is the basis of a portable yield strength tester presented by Granzow (G-31). The equipment is a hand-held bending device which bends the steel to a 90 degree angle (?) over a fixed radius R. A scale for a springback ?? is calibrated in strength for each of several common thicknesses of steel. Because of strain hardening, the yield stress is not measured. Instead a flow is measured which describes the stress state during the 90 degree bend. Granzow shows a correlation with the average of the yield and ultimate tensile strengths. This tester probably is the only forming operation which is accurately described by equation [1]. Work Hardening Effect - The equation of Gardner above [1] assumes no strain hardening, which is not realistic for most sheet metal forming for the automotive industry. This problem was addressed by Queener and DeAngelis (Q-1) who derived from plasticity theory an equation which incorporated the n and K values of the stress strain equation:

for R/t < 30. For n = O and K = , equation [3] reduces to equation [2] times a factor of 1.15. Equation [3] shows that an increase in n and/or K increases the amount of springback. The work hardening increases the flow stress, which in turn increases the recovery increment. The equation assumes uniform distribution of the strain around the entire bend. In reality, the work hardening may change the strain distribution and the fundamental behavior of the forming mode. A comparable example may be the effect of yield point elongation. The YPE tends to generate discrete bands of deformation called kinks. These kinks do not take the radius of the bending tool but tend to assume a much tighter radius (Figure 8.3-5). This reduction in R reduces the amount of springback. b. BENDING WITH TENSION More important than the work hardening is the effect of a tensile stress component added to the bending stresses. This deformation mode is common in automotive stampings. As the sheets are formed over punch and die radii, tensile restrain is introduced. These restraints are generated in the binder surfaces from draw beads inserted into the binders and from friction developed over the various components of the die. These tensile forces restrict the draw components within a complex stamping and encourage the stretch component. This restraint also is used to set the shape of the stamping. During bending the neutral axis shift is dependent upon the relative magnitudes of the bending stresses and the tensile stresses. It can be shown that this neutral axis shift is extremely important to springback. More specifically, a neutral axis shift outside of the metal sheet is favorable. A method of springback analysis was presented by Duncan and Bird (D-23). A bilinear stress-strain curve was assumed (Figure 8.3-6) with the following mathematical description of each zone.

In these equations, P is the plastic strain hardening modulus for linear hardening and is similar to the n value for parabolic hardening. When the sheet is fully plastic, the springback equation becomes:

When the sheet is fully plastic, the springback is small; E is usually two or three orders of magnitude greater than P. This analysis by Duncan and Bird (D-23) indicates that to obtain good shape control, the tension in the sheet should slightly exceed the yield stress in order to achieve a small but finite plastic strain throughout the sheet. This analysis requires that the radius R be large enough so that the assumption of fully plastic tensile stress across the thickness be met. Finally, this tension must be applied as a post tension that is, after the bending but before the unloading. Some operations can be designed with either a pretension or post tension. An analysis was made by Wenner (W-17) of the affect of pre-versus post-tension in shallow panels with large R/t ratios (shell membrane theory is applicable). His analysis included punch friction, work hardening, and strain rate hardening, but over a range of plastic strain up to three percent. Wenner verified the analysis of Duncan and Bird for post strain (G-6). One key result of Wenners analysis is shown in Figure 8.3-7 (W-17). He showed that springback was greater (smaller value of Shape Retention Factor) for pretension than post-tension. However, this result is overshadowed by the effect of the Tensile Load/Yield Load Ratio. As this ratio exceeds unity, (stamping is everywhere plastic), the Shape Retention Factor climbs rapidly. When the (applied) load is near the yield load, large differences in springback can occur for small differences in operating conditions. For the critical region where the tensile load approximates the yield load, the springback is very sensitive to changes in material properties (yield stress and thickness) and the forming process (restrain, alignment, fiction, etc.). Wenner concluded that in practice there may be a fine line between draw die settings that produce satisfactory parts and those that do not (W-17). Wenner continued the analysis to show that since the amount of springback was dependent on the amount of tension that could be developed over a punch with various amounts of friction, it was important to know if sufficient tension could be developed for particular steels (W-17) He showed that the tensile force developed depended on the ratio of the ultimate to yield strengths. He also showed that was very difficult to produce sufficient loads for dual-phase steel and HSLA-80 steel except for small angles of draw and/or low coefficients of friction. The AK steel, on the other hand, exhibited the lowest springback since a load greater than the yield value could be developed over a wide range of die angles and frictional values. c. CHANNEL BENDING Several studies have been conducted on the effect of a tensile load added to the channel bending operation. The combination has been characterized as a stretch-bending operation. The study by Neda and Veno treated the channel as a beam and calculated springback for a work hardening material (N-9). They showed that springback can be greatly reduced by applying tension. The tension was applied during and after the bending. One method of stabilizing springback in channel forming of materials with different yield stresses is to apply an optimum initial tension calculated from the initial yield stress often is unknown, a Fixes Initial Elongation method is used in which a constant elongation (such as one percent) is imparted to the material. This automatically compensates for the differences in yield stress. Springback in trapezoidal shaped corrugations and Z-shaped purling is discussed by Ingvarsson (I02). Again, the importance of stretch forming in reducing springback is cited.

SIDEWALL CURL Sidewall curl is a problem in channel forming which is superimposed on the traditional springback of sidewall opening (Figure 8.3-8). Sidewall curl is caused by friction. The friction between the sheet and the punch/die causes the tension to vary from surface to surface and along any given surface (Figure 8.3-9). Relaxation of these different tensions causes the curling action. A number of studies have dealt with sidewall curl and several press shop programs are designed to eliminate this sidewall curl (H-11). One such study by Vecchio provides side wall curl and tangent wall angles for 26 U-shaped automobile underbody rail specimens produced from mild steel, HSLA steel, and dual-phase steels formed with both flat punches and beaded punches (V-3). Two analyses have been completed by Wang (W-9). One analysis shows in Figure 8.3-10 that side wall curl can be reduced (eliminated) if enough side wall loading can be applied to cause sufficient stretching (W-9). This analysis by Wang (W-9) has enabled GM Research to develop a two-step process to eliminate sidewall curl. If the necessary tension is applied during forming in a single stage, a sever shock line and breakage would develop. Therefore, in the process called SHAPESET, the steel is the first bent to ahpe without draw beads and without tension; this creates severe sidewall curl (G-3, A-32). This sidewall curl is then removed in a second or stretch operation with draw beads (Figure 8.3-11). Other papers discussing sidewall curl include V-1, L-8, L-15, D-4, H-13. SHAPESET is an excellent example of where theory and experimentation have been used in tandem to provide a solution to a production forming problem.

8.3.4.2 Small Radius Bends Along Curved Lines


Springback in parts with contoured flanges and channels with sweep (for example frame rails and other structural parts) causes changes in both the bend angle (viewed in the cross-section) and elastic curvature of adjacent surfaces causing twist and sweep changes. At the present time, a comprehensive model for this problem does not exist (G-6). The problem with such flange or channel forming is that both the bend radius and the flange (or channel wall) are plastically deformed (Figure 8.3-12). A complex interaction of the recovery strains now results. A strong mechanical multiplying factor is present. A slight change in line length A-B can cause a large change in radius R and therefore cause a camber in the vertical surface.

8.3.4.3 Doubly-Curved, Shallow Smooth Panels


Springback in outer body panels and other doubly-curved, shallow, smooth panels depends on the distribution of both strain and residual stresses. Here again a complete model is lacking, but the problem can be easily illustrated (G-6). While the origin of springback in these shallow panels is different from small radius bends, a mechanical multiplying factor is involved. Small in-plane movement can cause a large deflection as illustrated in Figure 8.3-13. In some panels, a circular or hoop stress is created during the forming operation which causes a crown of height, h, to develop (Figure 8.3-14). The springback problem in shallow, outer body panels is best controlled by the introduction of a finite and uniform plastic strain; this is accomplished by controlled stretching of the sheet as shown in Figure 8.3-15. This stretching must introduce a tension in the sheet somewhat greater than the yield stress times thickness (G-6). The State of the Art Review by Gerdeen and Duncan contains a review of a number of papers which have some relationship to springback doubly-curved, shallow, smooth panels (G-6). Unfortunately, the results discussed often are contradictory to each other and, at best, bear only a slight relationship to real world panels. The problem seems to be that simple analyses are being applied to very complex stampings; the simplifying assumptions overwhelm the analysis so much that the analysis has no hope of describing panel behavior. Accurately predicting quantitative corrective action is even further beyond analysis. For this reason no review of these studies will be included here.

8.3.5 SUMMARY
The problem with springback can be summarized with the aid of Figure 8.3-16. This figure has been developed by a number of authors (G-6, J-5) and is similar to Figure 8.3-7. Most stampings simple bends or doubly-curved, shallow, smooth panels are formed such that the stretch level is in zone A. In this zone, the springback is sensitive to the type of metal (Youngs modulus), the thickness of the metal, the strength of the metal (yield stress), the strain imparted to the sheet, the uniformity of the strain, the residual stresses, the friction, and all other variables in the Forming System (K-23). When operating in this zone, the press operator correctly deduces the need for uniformity of thickness and mechanical properties of the sheet metal being used. He can not directly observe the variability in the other components of the Forming System. The evidence is clear both theoretical and experimental that imposing a stretch component to the stamping will either eliminate most of the springback or at least bring it to a constant value independent of material and process variables (zone B in Figure 8.3-16). What is needed are some well documented case histories which will illustrate this conclusion and provide guidelines for part designers and die designers alike. Figure 8.3-1 Unloaing from a plastic forming stress (A) results in springback or recovery strain (BC).

Figure 8.3-3 Diagram showing how a small movement is mechanically magnified into a large transverse movement in two cards joined only at one end (G-6). This experiment simulates a sheet

of metal during elastic recovery from a bending strain.

Figure 8.3-4 Springback defined as / for a bend ratio of R/t.

Figure 8.3-5 A sheet of metal may "kink" while being bend around radius R, thereby greatly

reducing the bend radius Rs. One common cause of kinking is yield point elongation.

Figure 8.3-6. A bilinear stress (-strain () curve.

Figure 8.3-7 Springback under pretension and post-tension as a function of the ratio of tensile

load to yield load (W-17).

Figure 8.3-8 Schematic showing sidewall opening and sidewall curl.

Figure 8.3-9 Different tensil stresses are developed at points C and D due to friction between the sheet and the punch/die. Elastic recovery of these different stresses causes sidewall curl.

Figure 8.3-10 Prediction of sidewall curl by the matematical model of Wang (W-09).

Figure 8.3-11 Schematic showing the two-stage "SHAPESET" process for elimination of sidewall

curl (G-3).

Figure 8.3-12 A contour flange (stretch-flange) of radius R1 formed from a flat sheet.

Figure 8.3-13 Demonstration in a flat sheet that small in -plane movements are mechanically magnified and can cause large transverse deflections (G-6).

A_1

Adak, I. and Travis, F.W., Optimization in the Pressing of ridges into Flat Panels, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 1721, 1978), pp. 67-78.

A_2

Ahrndt, G., Kugler, J., and Vlad, C. M., Some Observations on

A_3

A_4

A_5

A_6

A_7

A_8

A_9 A_10

A_11

A_12 A_13

A_14 A_15 A_16 A_17

the FLC Diagrammes of Direct Off the Rolling Heat Processed Dual Phase Steels, Proceedings of the Working Group Meetings of the IDDRG, S. Margherita Ligure, Italy, (May 24-28, 1982), Working Group Volume pp. 171-178. AIME, Computer Modeling of Sheet Metal Forming Processes: Theory, Verification, And Applications, Ann Arbor, Michigan, (April 1985), AIME, Warrendale, PA. Altan, T., Oh, S-1, and Gegel, H. L., Metal Forming Fundamentals and Applications, ASM, Metals Park, Ohio (1983). Altner, W., Description of the Micro-Surface of Sheet Metal Through Digital Image Processing, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 478-479. American Deep Drawing Research Group, Sheet Metal Forming Definitions, Published by American Society for Metals, Metals Park, Ohio, (1974). American Deep Drawing Research Group, Analysis of Sheet Metal Formability in Sheet Metal Forming and Formability, Proceedings of the 7th Biennial Congress of the International Deep Drawing Research Group, Amsterdam, (October 9-13, 1972), pp. 18.1 to 18.7. American Deep Drawing Research Group, Joint Meeting with the Japanese Deep Drawing Research Group, Ann Arbor, Michigan, October 12, 1976. Unpublished. American Deep Drawing Research Group, Sheet Metal Industries, 54, (1977), pp. 147-153. American Iron and Steel Institute, Automotive Steel Design Manual, AISI, Washington, D.C., (October 1986), pp. 4.1-10 to 4.1-18. American Iron and Steel Institute, Cost Effective Weight Reduction With Sheet Steel, AISI, Washington, D.C., (January 1977). American Iron and Steel Institute, Sheet Steel Formability, AISI, Washington, D.C., (1984). American Iron and Steel Institute, The Forming of Galvanized Sheet Steels Guidelines For Automotive Applications, AISI, Washington, D.D., (February 1986). American Society for Metals, Metals Handbook, 8th Edition, Vol. 4 (Mechanical Testing), ASM, Metals Park, Ohio, (1969). American Society for Metals, Metals Handbook, 9th Edition, Vol. 8 (Mechanical Testing), ASM, Metals Park, Ohio, (1985). American Society for Testing and Materials, Annual Book of ASTM Standards, ASTM, Philadelphia, PA. Antoniucci, P.L., V., and Spedaletti, M., A New Approach to

A_18

A_19

A_20

A_21

A_22

A_23

A_24

A_2

A_26 A_27

Surface Roughness Characterization, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 480-481. Aoki, I. and Horita, T., Development and Propagation of Surface Stain During Forming of Sheet Steel Panels, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 4 (April 1980), pp. 553-560. Arlinghaus, F. J., Frey, W. H., Stoughton, T. B., and Murty, B. K., Finite Element Modeling of a Stretch-Formed Part, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 5164. Arrieux, R., Bedrin, C., and Boivin, M., Determination of an Intrinsic Forming Limit Stress Diagram for Isotropic Metals Sheets, Proceedings of the Working Group Meetings of the IDDRG, S. Margherita Ligure, Italy, (May 24-28, 1982), Working Group Volume pp. 61-72. Arrieux, R., LeGac, H., and Sevestre, C., A New Experimental Method for the Determination of Forming Limit Diagrams at Necking, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 464-465. Arrigoni, G., Palladino, M., Volpato, G., and Gubian, S., Vacuum Alloyed and Decarburized HSLA Steels (HS VAD). FLC Obtained in Industrial Press of Styled Car Wheel Discs,: Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 4 (April 1980), pp. 485-493. Asai, T., Sakai, K., Kato, K., Oohashi, A., and Hirano, N., Study for the Use of High Strength Steel Sheet for Autobody Panels, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 3 (April 1980), pp. 257-266. Asai, T., et al, Improvement in the Forming Severity of Autobody Pressing, IDDRG Working Group I, Ann Arbor, MI, (1976). Asceirand, O. and Wilso, D. V., Effects of Microstructure on Formability in Biaxial Stretching, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 1721), 1978), pp. 155-156. ASTM, Test Method for Ball Deformation Test of Metallic Sheet Material, E643-78, Annual Book Of Standards. ASTM, Practice for Sheet Metal Forming Lubricant Evaluation, D4173-82, Annual Book Of Standards.

A_28 A_29

A_30

A_31 A_32

A_33

A_34 A_35

B_1

B_2 B_3 B_4 B_5

B_6

B_7

B_8

Atkinson, M. and Maclean, I. M., The Measurement of Plastic Anisotropy in Sheet Metal, IDDRG Coll., London (1964). Atkinson, M., Assessing Normal Plastic Anisotropy From Cup Drawing Loads, J. of the Iron & Steel Inst., 208, (January 1970) pp. 58-66. Ayres, R. A., Brazier, W. A., and Sajewski, V. F., Evaluating the GMR Limiting Dome Height Test as a New Measure of Press Formability Near Plane Strain, Journal of Applied Metalworking, 1, No. 1, (1979), pp. 41-49. Ayres, R. A. et al, J. Appl. Metalworking, 3, No. 3, (1984), p. 272. Ayres, R. A., Shapeset A Process to Reduce Sidewall Curl Springback in High Strength Steel Rails, J. of Applied Metalworking, 3, No. 2 (January 1984), pp. 127-134. Ayres, R. A., Thermal Gradients, Strain Rate and Ductility in Sheet Steel Tensile Specimens, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 436-443. Ayres, R. A., Technology Transfer: Making It Really Happen, S'E Paper No. 860127, (1986). Azrin, M. and Backofen, W. A., The Deformaiton and Failure of a Biaxially Stretched Sheet,: Met. Trans., 1, (1970), pp. 28572865. Babcock, S. G., High Speed Material Property Testing of Sheet Metal, Paper presented to IDDRG Working Group III, Amsterdam, (1972). Backofen, W. A., Deformation Processing, Addison-Wesley, Reading, Mass., (1972). Backofen, W. A., Private communication: Backofen, W., A., Massachusetts Institute of Technology Industrial Liaison Symposium, Chicago, April 1974. Barata da Rocha, A., Barlat, F., and Luken, P. C., Forming Limit of Anisotropic Sheets in Complex Strain Paths, Computer Modeling Of Sheet Metal Forming Process; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 93-106. Baret, J. C. and Wybo, M., LaCourbe Limite De Formage A LApparition De La Striction, CRM Report No. RA 838/71, July 1971, Presented at the BDDRG Meeting, London, 1971. Bath, C. J. and Mackenzie, R. A., Stretch Forming, 3rd International Conference of Sheet Metal Working, Chicago, (April 14, 1969). Bauer, M., Determining Flow Curves for Sheet Metal by Plane Torsion Tests, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 468-469.

B_9

B_10

B_11

B_12

B_13 B_14

B_15

B_16

B_17

B_17

Berland, J., Detraux, J. M., Horkay, F., Jalinier, J. M., and Manh, C., The Simulation of the Sheet Metal Forming Process Using Numerical Models of Increasing Complexity, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 341-352. Blickwede, D. J., New Knowledge About Sheet Steel, Metals Progress. Dec. 68 pp. 64-70 Jan. 69 pp. 84-87 Feb. 69 pp. 83-85 Apr. 69 pp. 87-92 May 69 pp. 86-88 June 69 pp. 120-122 July 69 pp. 91-94 Sept. 69 pp. 99-100 Nov. 69 pp. 77-80 Dec. 69 pp. 98-102 Jan. 70 pp. 86-89 Blumel, K., Surface Interactions During Pressforming as Simulated by the Strip Draw Test,: Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 1721, 1978), pp. 297-302. Boccalari, E., Daglio, R., Irmici, W., and Motano, G., The Use of HSLA Steel Sheets in Two-wheeled and Three-wheeled Vehicles, Proceedings of the 12th Biennial Congress of the IDDRG, S. Margherita Ligure, Italy, (May 24-48, 1982), pp. 2132. Bowden, F. P. and Tabor, D., Friction: An Introduction To Tribology, Heinemann, London, (1974). Brammer, I. S. and Harris, D. A., Production and Properties of Sheet Steel and Aluminum Alloys for Forming Application, Journal of the Austrialian Institute for Metals, 20, No.2, (1975), pp. 85-100. Brazier, W. G., The Application and Potential for Practical Models of Sheet Metal Formability, Proceedings of the 13th Biennial Congress of the International Deep Drawing Research Group, Melbourne, Australia, (February 1984). Brazier, W. G. and Thompson, R. W., The Effect of Die Materials and Lubrication on punch Stretching and Drawing Forces of Cold-Rolled and Zinc-Coated Steels, S'E Paper 850274, (1985). Brazier, W. G. and Thompson, R. W., The Effect of Zinc Coatings, Die Materials, and Lubricants on Sheet Metal Formability, S'E Paper 860434, (1986). Brazier, W. G., Die Design for High Strength to Weight Sheet

B_18

B_19 B_20 B_21

C_1

C_2

C_3 C_4

C_5

C_6

C_7 C_8

C_9

Materials, ASM/ADDRG Symposium on Application of High Strength-to-Weight Metals for Automotive Components, Troy, MI, (November 12, 1974). Brozzo, P. and DeLuca, B., On the Interpretation of the Formability Limits of Metal Sheets and Their Evaluation by Means of Elementary Tests, Proceedings of the International Conference on the Science and Technology of Iron and Steel, Tokyo, (September, 1970), pp. 966-968. Bucher, J. R. and Hambury, E. G., High Strength Formable Sheet Steel, S'E Paper No. 770164, (1977). Budiansky, D. and Wang, N. M., On the Swift Cup Test, J. of Mech. of Physics and Solids, 14, (1966), p. 357. Burst, H., Experiences at Porsche with Galvanized Steel, Detroit ASM Conference on Solutions to Manufacturing Problems Using Galvanized Steels, Summary by W. J. Riffe, AISI, Washington, D.C., (1985), pp. 1-4. Carlson, R. F., Basic Design Requirements for Stampings, American Metal Stamping Association Yearbook, (1968), pp. 50123. Chakhari, M. L. and Jalinier, J. M., Spring Back of Complex Bent Parts, Proceedings of the 13th Biennial Congress, Melbourne, Australia, (February 1984), pp. 148-159. Charpentier, P. L., Met. Trans., 6A, (1975), p. 1665. Charpentier, P. L. and Piehler, H. R., In-Plane Compression; A New Laboratory Forming technique to Simulate Drawing of Sheets, Novel Techniques In Metal Deformation, R. H. Wagoner-ed., TMS of AIME, Warrendale, PA, (1083), pp. 131148. Charpentier, P. L. and Piehler, H. R., Properties of Formed Parts: Flow and Fracture Behavior After Cold Forming in HSLA Steel Sheets, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 3 (April 1980), pp. 281-292. Chatfield, D. A. and Beiser, A. A., A Study of the Surface Irregularities Caused by the Nonhomogenious Nature of Temper Rolling, AISI Yearbook, (1968), pp. 163-185. Also Blast Furnace and Steel Plant, 56, (August 1968), pp. 696-727, 742. Chatfield, D. A. and Keeler, S. P., Designing for Formability, Metal Progress, 99, (1971), pp. 60-65. Chatfield, D. A. and Rote, R. R., Strain Rate Effects on the Properties of High Strength, Low Alloy Steels, S'E Paper No. 740177, (1974). Chu, E. and Sowerby, R., Some Aspects of Finite Deformation, Advanced Technology of Plasticity, Proceedings of the First International Conference of Technology of Plasticity,

C_10

C_11

C_12

C_13

C_14

C_15

C_16 D_1 D_2 D_3 D_4

D_5

D_-6

D_7

Tokyo, (1984), pp. 1065-1069. Chu, E., The Coaxial and Non-Coaxial Strain Paths in Finite Deformation, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 353-362. Chu, E., Soper, D., Gloeckl, H., and Gerbeen, J.C., Computeraided Geometric Simulation of Sheet Metal Forming Processes, Computer Modeling of Sheet Metal Forming Processes; Theory, Verification, And Application - Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 6576. Chu, E., Sowerby, R., Soldaat, R., and Duncan, R.L., Strain Measurement Over Large Areas of an Industrial Stamping, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 9-22. Chung, K. and Lee, D., Computer-aided Analysis of Sheet Material Forming Processes, Advanced Technology Of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 66-665. Chung, S.Y. and Swift, H.W., Cup-drawing From a Flat Blank: Part I. Experimental Investigation; Part II. Analytical Investigation, Proc. Inst. Mech. Engineers, 165, (1951), pp. 199. Cockerill, S. A., Compared Deep-Drawing Behavior of Mild Steel and HSLA Steel Cold Rolled Sheets, IDDRG Working Group II, Ann Arbor, MI, (1976). Crandall, S. H., Discussions on Gardners Paper, ASME Transactions, 79, 1, (1957), pp. 7-8. Danckert, J. and Wanheim, T., J. of Mech. Working Technology, 3, (1979), p.5. Darlington, H., Private Communication. Davies, R. G., Side-wall Curl in High-Strength Steels, J. of Applied Metalworking, 3, No. 2, (January, 84), pp. 120-126. Davies, R. G., and Liu, Y. C., Control of Springback in Flanging Operation, J. of Applied Metalworking, 3, No. 2, (January 1984), pp. 142-147. Davies, R. G., Chu, C. C., and Liu, Y. C., Recent Progress on the Understanding of Springback, Computing Modeling of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 259-271. Davison, R. M., Evaluation of High-Strength Low-Alloy Steel Sheets by Hole Enlargement Tests, Climax Moybdenum Company Report No. L-176-131, (May 24, 1974 Day, J. M., Flexible Panel Production, Proceedings of the 13th Biennial Congress of the International Deep Drawing Research

D_8

D_9

D_10

D_11 D_12

D_13

D_14 D_15

D_16

D_17 D_18

D_19

Group, Melbourne, Australia, (February 1984), pp. 288. deGroot, M., On the Interpretation of Results of Hole Expansion Tests on Sheet Steel, Sheet Metal Forming And Energy Conservation, Proceedings of the 9th Biennial Congress of the International Deep Drawing Research Group, Ann Arbor, MI (October 1976), pp. 193-206. Deleon, Y., Daigne, J., Lefevre, G., and El Haik, R., Sheet Metal Forming in the Automobile Industry. Respective Roles of the Deformation Capacity & Sliding of the Sheet Metal on the Forming Tools, Proceedings of the 12th Biennial Congress of the IDDRG, S. Margherita Ligure, Italy, (May 24-28, 1982), pp. 355-366. DeLuca, B., Industrial Applications of Formability Limit Curves and A First Approach to Strain Paths, Paper Presented to the IDDRG Working Group I, Zurich, September, 1973. Demeri, M. Y., The Stretch-Bend Forming of Sheet Metal, J. of Applied Metalworking, 2, No. 1, (1981), pp. 3-10. deSouza Nobrega, M. C., deAndrale, H. L., Sarmento, E. C. and Ferran, G., Evolution of the Microscopical Heterogenities in Sheet Metal Forming, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17-21, 1978), pp. 145-154. evenpeck, M. C. and Richmond, O., Limiting Strain Tests for In-Plane Sheet Stretching, Novel Techiques In Metal Deformation, R. J. Wagoner ed., TMS of AIME, Warrendale, PA, (1983), pp. 79-88. Dieter, G. E. (editor), Workability Testing Techniques, ASM, Metals Park, Ohio, (1984). Dieter, G. E., Tension Testing, Workability Testing Techniques, G. D. Dieter ed., ASM, Metals Park, Ohio (1984), pp. 21-36. Dieter, G. E., Strain-Rate Effects in Deformation Processing, Fundamentals Of Deformation Processing, Proceedings of 9th Sagamore Conference, Syracuse University Press, (1964), pp. 145-182. Dinda, S. et al, High-Strength Steel for Cost Effective Weight Reduction, S'E Paper No. 760207, (1976). Dinda, S., James, K. F., Keeler, S. P., and Stine, P. A., How To Use Circle Grid Analysis For Die Tryout, American Society for Metals, Metals Park, Ohio, (1981). Doege, E., Prediction of Failure in Deep Drawing, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., the Metallurgical Society of AIME, Warrendale, PA (1986), pp. 209224.

D_20

D_21 D_22 D_23

D_24 D_25

E_1 E_2

E_3 E_4

E_5

E_6

F_1

F_2

F_3

Doege, E. and Witthuser, K. P., HFF-ERICHTS, UKH, Gessellschaft fur Produktionstechnik, Hannover, Germany, (1977), pp. 129-137. Duncan, J. L. and Altan, T., New Directions in Sheet Metal Forming, CIRP Annals, 29, No. 1, (1980), p. 153. Duncan, J. L., Shabel, B. S., and Filho, J. G., S'E Paper No. 780391, (1978). Duncan, J. L.. and Bird J. E., Die Forming Approximations for Aluminum Sheet, Sheet Metal Industries, September (1978), pp. 1015-1023. Duncan, J. L. and Sowerby, R., Computer Aids in Sheet Metal Forming, CIRP ANNALS, (30), No. 2, (1981), pp. 541-544. Duncan, J. L., Sowerby, R., and Chu, E., The Development of Sheet Metal Modeling (Keynote Paper), Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 1-11. Eary, D., Class Notes, Unpublished. Eary, D. F. and Reed, E. A., Techniques Of Pressworking Sheet Metal, 2nd Ed., Prentice-Hall, Englewood Cliffs, New Jersey, (1974). Ebben, G. J., American Can, Private Communication. Ejima, M., Tokunaga, Y., and Honda, T., Analysis of Formability and Some Problems in Stamping of Coated Steel Sheets, Proceedings of the 13th Biennial Congress of the International Deep Drawing Research Group, Melbourne, Australia, (February 1984), pp. 317-328. El-Sabaie, M. G. and Mellor, P. B., Prediction of the Limiting Drawing Ratio in the Deep Drawing of a Flat-Headed Cylinder Shell, in Sheet Metal Forming And Formability, Proceedings of the 7th Biennial Congress of the International Deep Drawing Research Group, Amsterdam, (October 9-13, 1972), pp. 2.1 to 2.5. Emmens, W. C. and Hartman, L., The Effect of Surface Roughness on the Formability of Cold-Reduced Sheet Steel, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 491-504. Fernandez, J., Riba, J., and Verdeja, J. I., Damage and Its Evolution in a Microalloyed Steel Sheet, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 414-415. Ferry, B. N. and Paliwoda, E. J., The Forming of Vanadium Bearing HSLA Steels Into Automotive Components, S'E Paper No. 740180, (1974). Ferry, J. W., Techniques and Limitations for Cold Roll Forming

F_4

F_5

F_6

F_7

F_8

F_9

F_10

F_11

F_12

G_1 G_2 G_3 G_4 G_5

High Strength Steel Bumper Components, S'E Paper, No. 810120, (1981). Fine, T. E. et al, Evaluation of Edge Formability Properties of Hot Rolled Sheet Steels, Mechanical Working And Steelmaking, Metallurgical Society Conferences, Pittsburgh, PA, (January 1977), pp. 1-31. Fogg, B., Modern Developments in Lubrication Theory and Practice for Deep Drawing, Sheet Metal Industries, 53 (May 1976), pp. 294-304. Frommann, K. M., The Influence of Strain Aging on the Formability Characteristics of Rimmed Steel, Republic Steel Research Report (November 30, 1967), Publication Unknown. Frommann, K. M., The Influence of Strain Rate and Punch Speed on the Mechanical Properties and the Drawability of Deep Drawing Steels, Republic Research Report, (September 11, 1968), Publication Unknown. Fromman, K. M., The Prediction of Metal Stamping Behavior, Amer. Metal Stamping Assoc. Spring Tech. Conf., Detroit, MI (April 17, 1968). Fukui, S., Conical Cup Test Research Group, Institute of Physical and Chemical Research, Komagome, Bunkyo-Ku, Tokyo, (December 1, 1958). Furubayshi, T., An Investigation Into High Strength Steel Sheets and Aluminum Alloy Sheets for Press Forming of Autobody Parts, IDDRG Working Group I, Ann Arbor, MI, (1976). Furubayashi, T., Sato, S., Hirasaka, M., and Yoshihara, N., Technological Impact Of Surfaces Relationship To Forming, Welding, And Painting, American Society for Metals, Metals Park, Ohio (1982), pp. 147-160. Furubayshi, T., Shiokawa, M., and Yamazaki, K., Press Performance, Forming Geometry and Fracture Behavior in Actual Autobody Panels, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17-21, 1978), pp. 39-44. Gardiner, F. J., The Springback of Metals, ASME Transactions, 79, 1, (1957), pp. 1-6. General Motors Research Laboratory, Mathematical Modeling Improves Sheet Metal Stampings, Search, 12, No. 4, (1977). General Motors Research Laboratory, Showing Sheet Steel Whos Boss, Search, GM Research Publication, 18, 2, (1983). George, R. A. et al, Estimating Yield Strength From Hardness Data, Metal Progress, 109, (May 1976), pp. 30-35. Gerdeen, J. C., Bhonsle, S. R., and Wong, V., Stress Analysis of Yoshida Buckling Test Specimen Using Photoelasticity and

G_6 G_7

G_8

G_9

G_10

G_11

G_12

G_13

G_14

G_15 G_16

G_17

Finite Element Method, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 476-477. Gerdeen J. C. and Duncan, J. L., Springback in Sheet Metal Forming, Report on AISI Project 1201-456, (1986). Gerdeen, J. C. and Brazier, W. G., The LDH Test Experiments and Computer Simulation, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 315-328. Ghosh, A. and Backofen, W. A., Strain Hardening and Instability in Biaxially Stretched Sheets, Met. Trans., 4, (1973), pp. 1113-1123. Ghosh, A. K., The Effect of Lateral Drawing-in on Stretch Formability, Metal Engineering Quarterly, 15, (August 1975), pp. 54-64. Ghosh, A. K., The Influence of Strain Hardening and Strain Rate Sensitivity on Sheet Metal Forming, Journal of Engineering Materials and Technology (Trans. ASME), 99, (1977), pp. 264-274. Ghosh, A. K. and Hamilton, C. H., Superplastic Forming of Long Rectangular Box Section Analysis and Experiment, Process Modeling Fundamentals And Applications To Metals, American Society for Metals, Metals Park, Ohio, (1980), pp. 303-331. Ghosh, A. K. and Hecker, S. S., Stretching Limits in Sheet Metals: In-Plane vs. Out-of-Plane Deformations, Met. Trans., 5, (1974), pp. 2161-2164. Ghosh, A. K., Hecker, S. S., and Keeler, S. P., Sheet Metal Forming and Testing, Workability Testing Techniques, G. D. Dieter ed., ASM, Metals Park, Ohio, (1984), pp. 135-196. Ghosh, A. K., Plastic Properties in Relation to Localized Necking in Sheets, Mechanics Of Sheet Metal FormingMaterial Behavior And Deformation Analysis, General Motors Research Symposium, Plenum Press, New York, 1978, pp. 287311. Ghosh, A. K., How to Rate Stretch Formability of Sheet Metals, Metal Progress, 107, (May 1975), pp. 52-53. Ghosh, A. K. and Laukonis, J. V., The Influence of Strain-Path Changes on the Formability of Sheet Steel, Sheet Metal Forming And Energy Conservation, Proceedings of the 9th Biennial Congress of the International Deep Drawing Research Group, Ann Arbor, MI, (October 13-14, 1976), pp. 167-178. Ghosh, S. K. and Travis, F. W., Deformation Behavior of Then Diaphragms in Static and Dynamic Punch-Stretching, Proceedings of the 10th Biennial Congress of the IDDRG,

G_18

G_19

G_20

G_21

G_22 G_23

G_24 G_25

G_26

G_27 G_28 G_29

G_30

G_31

G_32

Warwick, England (April 17-21, 1978), pp. 95-112. Gibson, T. J., Hobbs, R. M., and Stewart, P. D., in International Conference Product Technology, Institution of Engineers, Melbourne, Australia, (1974), pp. 328-332. Gierzynska, Soln, M. and Bociaga, E., Tribological Attempts to Estimation of Seizing Ability of Metal Sheets, Advanced Technology of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 157161. Glauzou, J., Theoretical Basis of Forming Limit Diagram, Paper presented to IDDRG Working Group I, Zurich, (September 1973). Godard, F. and Renard, L., Everyday Use of an Automatic Tensile Testing Machine for Routine Control of Steel Sheets, IDDRG Working Group III, Ann Arbor, MI, (1976). Goodhart, R. R., Private Communication. Goodwin, G. M., Application of Strain Analysis to Sheet Metal Forming Problems in the Press Shop, S'E Paper No. 680093, (January 1968). Goodwin, G. M., Formability Index, SME Paper No. MF71165, (1971 Goto, Y., Wakasugi, S., and Furukawa, T., A Study of Material Pickup to Tool Surface in Metal Forming, Advanced Technology Of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 162167. Gotoh, M., A New Plastic Constitutive Equation and Its Applications to Forming Limit Problems of Metal Sheets, Advanced Technology of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 695-700. Gradous, F. J., Roll Form Design II Section Orientation, Tooling and Production, September 1966. Gradous, F. J., Roll Form Design III Locating the Vertical Guideline, Tooling and Production, November 1966. Gradous, F. J., Roll Form Design V Strip Width Calculations and the Section Flower, Tooling and Production, February 1968. Gradous, F. J., Roll Form Design VII Roll Pass Design Placement of Bend Allowance, Tooling and Production, (Aug. 1969). Granzow, W. G., A Portable Springback Tester for In-Plant Determination of the Strength of Sheet Steels, S'E Paper No. 830238 (1983). Granzow, W. G., The Effect of Tooling Temperature on the

G_33

G_34

G_35

G_36

G_37

G_38

G_39

G_40

G_41

G_42

Formability of Low Carbon Steel Sheets, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17-21, 1978), pp. 127-132. Green, S. J. et al, Material Properties, Including Strain Rate Effects, As Related to Sheet Metal Forming, Met. Trans., 2, (1971), pp. 1813-1820. Green, A. P., Friction Between Unlubricated Metals A Theoretical Analysis of the Junctions Model, Proc. Royal Society, A, 208 (1953), pp. 191-204. Gronostajski, J. and Dolny, A., Determination of Forming Limit Curves by Means of Marciniak Punch, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 3 (March 1980) and No. 4 (April 1980), pp. 570-578. Gronostajski, J., Use of New Yield Criteria in Determining the Effect of Strain Path on Forming Limit Diagrams, Advanced Technology of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 723727. Gronostajski, J., Dolny, A., and Sobis, T., Formability of High Strength Steel and ARMCO Iron at Different Strain Path, Proceedings of the Working Group Meetings of the IDDRG, S. Margherita Ligure, Italy, (May 24-28, 1982), Working Group Volume pp. 39-48. Grononstajski, J., Ghattas, M. S., and Ali, W. J., The Effect of Zinc and Plastic Coating on the Formability of Steel Sheet, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 454-455. Grumbach, M. and Sanz, G., Influences de Grilles de Mesure des Trajectories de Deformation sur les Courbes Limites d Emboutissage, in Sheet Metal Forming And Formability, Proceedings of the 7th Biennial Congress of the International Deep Drawing Research Group, Amsterdam, (October 9-13, 1972), pp. 4.1 to 4.7. Grumbach, M. et al, Quantitative Study of the Relationship Between the Strain Ratio (r value), Earring, and the Texture of Thin Sheets Low Carbon Steel, Proceedings of the 8th Biennial Congress of the International Deep Drawing Research Group, Gothenburg, Sweden, (September 23, 1974), pp. 62-76. Grzeski, D. and Vlad, C. M., Influence of Deformation Path Upon the Change of Plastic Anisotropy During the Forming of Hot and Cold Rolled Low Carbon Deep Drawing Sheet Steel, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17-21, 1978), pp. 203-210. Gutowski, T. A., A Logical Approach to Roll Forming

G_43

H_1 H_2 H_3 H_4

H_5

H_6 H_7

H_8

H_9

H_10

H_11

H_12

Tooling, SME/FMA FABTECH International Conference, September 1981. Gutowski, T. A., Cold Rollforming: An Introduction, SME/FMA FABTECH International Conference, 1983, pp. 3-1 to 3.9. Haberfield, A. B., and Boyles, M. W., Sheet Metal Industries, 50, (1973), p. 400. Halmos, G. T., Length Tolerances for Roll Formed Parts, Precision Metal, December 1981. Halmos, G. T., Design for Manufacturability, High Production Roll Forming, SME, Dearborn, MI, 1982, pp. 192-211. Harlifinger, R., Jurgetz, A., and Steidl, M., Surface Quality Requirements for Sheet Metal in Car Bodies, Proceedings of the 14th Biennial Congress of the International Deep Drawing Research Group, Koln, Germany, (April 1986), pp. 219-230. Harvey, D. N. Optimizing Patterns and Computational Algorithms for Automated, Optical Strain Measurement in Sheet Metal, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 403414. Hasek, V., Ind. Anz., 99, (1977), p. 343. Havranek, J., The Effect of Mechanical Properties of Sheet Steels on the Wrinkling Behavior During Deep Drawing of Conical Sheels, Sheet Metal Forming And Energy Conservation, Proceedings of the 9th Biennial Congress of the International Deep Drawing Research Group, Ann Arbor, MI, (October 13-14, 1976), pp. 245-263. Havranek, J., Wrinkling Limit of Tapered Pressings, Journal of the Australian Institute for Metals, 20, No. 2, (1975), pp. 114119. Hawtin, L. R., Recommended Procedure for Performing the Swift Cup Test, Sheet Metal Industries, 46, (May 1969), pp. 418-421. Hayashi, Y., The Study on Resistance Against Wall Breakage of Metal Sheets, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17-21, 1978), pp. 87-94. Hayashi, Y., Analysis of Surface Defects and Side Wall Curl in Press Forming, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 2o-25, 1984), pp. 565580. Hayashi, Y., Shibasaka, O., and Yoshida, K., Deforming Behavior of a Triangular Specimen in the Diagonal Tensile Test, Proceedings of the Working Group Meetings of the International Deep Drawing Research Group in S. Margherita Ligure, Italy, (May 1982), pp. 179-190.

H_13

H_14 H_15 H_16

H_17 H_18 H_19

H_20

H_21

H_22

H_23

H_24 H_25

H_26

H_27

Hayashi, Y. and Takagi, M., Control of Sidewall Curl in Draw Bending of High Strength Steel Sheets, Advanced Technology Of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 735-740. Heck, S. S. and Ghosh, A. K., The Forming of Sheet Metal, Scientific American, 235, (1976), pp. 100-108. Hecker, S. S. and Ghosh, A. K., The Forming of Sheet Metal, Scientific American, 235, (1976), pp. 100-108. Hecker, S. S., Formability of Aluminum Alloy Sheets, Journal of Engineering Materials and Technology, 97, (January 1975), pp. 66-73. Hecker, S. S., Formability of High-Strength Low-Alloy Sheets, Met. Eng. Quarterly, 13, No. 3, (1973), pp. 42-48. Hecker, S. S., A Cup Test for Assessing Stretchability, Met. Eng. Quart., 14, (November 1974), p. 30-36. Hecker, S. S., A Simple Forming Limit Curve Technique and Results on Aluminum Alloys in Sheet Metal Forming And Formability, Proceedings of the 7th Biennial Congress of the International Deep Drawing Research Group, Amsterdam, (October 9-13, 1972), pp. 5.1 to 5.8. Heyer, R. H. and Newby, J. R., Measurement of Ductility of Low Carbon Steel Sheets, Mechanical Working And Steel Processing V, Metallurgical Society Conferences, Vol. 50, (1966), pp. 133-154. Heyer, R. H. and Newby, J. R., Measurement of Strain Hardening and Plastic Strain Ratio Using the Circle-Arc Specimen, Sheet Metal Industries, 43, (1966), pp. 910-914. Heyer, R. H. and Newby, J. R., Effects of Mechanical Properties on Biaxial Stretchability of Low Carbon Steels, Trans. S'E, 77, (1969), Also S'E Paper No. 680094. Heyer, R. H., et al, Grain Orientation of Drawing Quality Steel Sheets, Flat-Rolled Products-III, Metallurgical Society Conferences Vol. 16, Interscience Publishers, 1962, pp. 29-46. Hiam, J. R. and Lee, A., Sheet Metal Industries, 55, (1978), p. 631. Hiam, J. R., The Effect of Sulphide Inclusions on Cold Formability, ASM Conference on Sulphide Inclusions in Steel, November 7, 1974. Higashi, M., Mori, T., Taniguchi, H., and Yoshimi, N., Geometric Modeling for Efficient Evaluation of Press Forming Severity, Computer Modeling of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 21-35. Hilsen, R. R. and Bernick, L. M., Relationship Between

H_28

H_29

H_30

H_31

H_32 H_33

H_34

H_35

H_36

H_37

H_38 H_39 H_40 H_41

Surface Characteristics and Galling Index of Sheet Steel, ASTM Formability 77, Toronto, Ontario, (May 4, 1977), pp. 129. Hilsen, R. R. et al, "Stamping Potential of Hot-Rolled. Columbium-Bearing High-Strength Steels," Proceedings Of Microalloying 75, (1977), pp. 654-664. Hilsen, R. R. et al, Stamping Potential of Hot-Rolled. Columbium-Bearing High-Strength Steels, Proceedings Of Microalloying 75, (1977), pp. 654-664 Hilsen, R. R., Surface and Galling Characteristics of a Series of High Strength Cold Rolled Steels, Working Group Meeting, International Deep Drawing Research Group, Warwick, England, April 1978. Hira, T., Azuma, S., Abe, T., Sasaki, T., and Yoshida, K. Analysis of Strain Propagation Behaviour in Yoshida Buckling Test, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 543553. Hobbs, R. M., Sheet Metal Industries, 55, (1978), p. 451. Hobbs, R. M., Pressing Automobile Body Panels Methods to Assess Forming Severity, BHP Tech. Bull., 25, No. 2, (Nov. 1981), pp. 49-59. Hobbs, R. M., Prediction and Analysis of Press Performance for Sheet Steels, BPH Tech. Bull. 18, No. 2 (November 1974), pp. 2-14. Hobbs, R. M. et al, Sheet Metal Forming Transferring Technology to the Shop Floor, Sheet Metal Forming And Formability, Proceedings of the 10th Biennial Congress of the International Deep Drawing Research Group, Warwick, England, (April 1978), pp. 121-126. Hochon, B., Deleon, Y., Jussiaume, A., and Perry, D., Methods of Evaluating the Fabricability of Precoated Sheet Metal and Repercussions on the Quality of Car Components, Proceedings of the 14th Biennial Congress of the International Deep Drawing Research Group, Koln, Germany, (April 1986), pp. 239-250. Hosford, W. F., The Effect of Anisotropy and Work Hardening on Cup Drawing, Redrawing, and Ironing, (1977), Publication Unknown. How Many Passes?, Tooling and Production, June 1966. Hu, H., The Strain Dependence of r Value of Textured Steel Sheet, IDDRG WG III, Ann Arbor, MI, (1976). Hugo, H. R., Good Stampings Start with Dies, Iron Age, 203, (May 22, 1969), pp. 72-74. Hugo, H. R., How to Improve Metal Stamping Die Performance, Metal Stamping, (October 1969).

H_42

H_43

H_44

H_45

H_46

I_1 I_2

I_3 I_4 I_5

I_6

I_7

I_8

Hultgren, F. A. and Barclay, W. F., Factors Influencing the Deep Drawing Quality of Rimmed Steel, Republic Steel Research Center Report, (July 26, 1966), Publication Unknown. Humayun, A., Roll Forming of Metal Coated Sheets, Presented at the Annual Meeting of the European Coil Coating Association, Venice, Italy, June 1986. Hutchinson, J. W., Neale, K. W., and Needleman, A., Sheet Necking I. Validity of Plane Stress Assumptions of the LongWave Length Approximation, Mechanics Of Sheet Metal Forming Material Behavior And Deformation Analysis, General Motors Research Symposium, Plenum Press, New York, (1978), pp. 111-126. Hutchinson, J. W. and Neale, K. W., Sheet Necking II. TimeIndependent Behavior, Mechanics Of Sheet Metal FormingMaterial Behavior And Deformation Analysis, General Motors Research Symposium, Plenum Press, New York, (1978), pp. 127-150. Huthinson, J. W. and Neale, K. W., Sheet Necking III. Strain Rate Effects, Mechanics Of Sheet Metal Forming Material Behavior And Deformation Analysis, General Motors Research Symposium, Plenum Press, New York, (1978), pp. 269-283. Ike, H., Yoshida, K., and Murakawa, Wear, 72, (1981), pp. 143155. Ingvarsson, L., Cold Formed Residual Stresses in Thin Walled Structures, Int. Congress on Thin-Walled Structures (Univ. of Stratchclydie), (1979), pp. 575. International Deep Drawing Research Group, Report of FLD Sub-Committee, Zurich, (1973). International Organization for Standardization, Recommendation R87. Iseki, H., and Murota, T. Analysis of Deep Drawing of NonAxisymmetric Cup by the Finite Element Method, Advanced Technology Of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 678684. Ishigaki, H., Umehara Y., and Okamoto, I., Analysis of Surface Deformation Defects in the Press Forming of Large Sized Autobody Panels, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17-21, 1978), pp. 53-66. Ishigaki, H., Deformation Analysis of Large Sized Panels in the Press Shops, Mechanics Of Sheet Metal Forming Material Behavior And Deformation Analysis, General Motors Research Symposium, Plenum Press, New York, (1978), pp. 315-338. Ivaska, Analysis of Lubrication Problems in Roll Forming, SME/FMA FABTECH International Conference, September

I_9 J_1

J_2

J_3

J_4

J_5 J_6 J_-7

J_8 J_9

J_10 J_11

J_12

J_13

1981, pp. 256-268. Ivaska, N., A General Survey of Sheet Surface Damages in Press Shops, IDDRG WG I, Ann Arbor, MI (1976). Jacobi, W., Development of Press-Forming Processes for Autobody Panels, Proceedings of the 14th Biennial Congress of the International Deep Drawing Research Group, Koln, Germany, (April 1986), pp. 31-62. Jalinier, J. M., Schmitt, J. H, Argemi, R., Salsmann, F. L., and Baudelet, B., Different Damage Behaviors and Their Influence on Forming Process, Proceedings of the 11th Biennial Congress of the IDDRG, Memories Scientifiques Revue Metallurgie, No. 3 (March 1980), pp. 313-327. Jalinier, J. M. and Baudelet, B., Theoretical Analysis of the Influence of Damage on the Shape and Position of the Forming Limit Diagram and on the Strain-Path and Thickness Effects, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17-21, 1978), pp. 133-144. Jalinier, J. M., Hayashi, H., Shibaski, O., and Yoshida, L., Analysis of Elastic Recovery by Means of the Yoshida Buckling Test, Proceedings of the 12th Biennial Congress of the International Deep Drawing Research Group in S. Marghertia Ligure, Italy, (May 1982), pp. 201-210. James, K. F., Unpublished work. James, K. F., Factors Affecting the Formability of Steel, GM Manuf. Dev. Report, (December 1970), Publication Unknown. The Japan Sheet Metal Forming Research Group, Galling Behavior and its Tests in Press Forming of Autobody Parts, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17-21, 1978), pp. 287-296. Jevons, J. O., The Metallurgy Of Deep Drawing And Pressing, John Wiley and Sons, New York, (1940). Johannison, T. G., High Pressure Flexible Die Forming of Sheet Metal for the Automotive Industry, Proceedings of the 14th Biennial Congress of the International Deep Drawing Research Group, Koln, Germany, (April 1986), pp. John Lysaght, (Australia), Sheet Steel Fabrication Handbook 3: Pressing And Forming, John Lysaght, Sydney, Australia. Johnson, W. and Duncan, J. L., The Use of the Biaxial Test Extensometer, Sheet Metal Industries, 42, (April 1965), pp. 271-275. Johnson, W., Ghosh, S. K., and Reid, S. R., Piercing and Holeflanging of Sheet Metals: A Survey, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Meallurgie, No. 4 (April 1980), pp. 585-606. Johnson, W. E., Stamping, Blanking, and Drawing, ASM

J_14

J_15 J_16

K_1

K_2

K_3

K_4 K_5 K_6

K_7

K_8

K_9

K_10 K_11

Conference on Cold Forming of Metals, Boston, (March 1970). Johnson, W. E. and Stine, P. A., Application of Sheet Metal Research to High Production Fabrication Shops, Sheet Metal Forming And Energy Conservation, Proceedings of the 9th Biennial Congress of the International Deep Drawing Research Group, Ann Arbor, MI, (October 13-14, 1976), pp. 235-244. Jorgens, L., Computerized Tooling for Roll Forming, Precision Metal, (June August, 1977). Jun, G-C, Effect of Changing Load Path on Yielding and Flow Behavior of Sheet Steels, Doctoral Thesis, Metallurgical Engineering Department, University of Michigan, Ann Arbor, Michigan, (1984). Kaftanoglu, B. and Kilkis, B., Theory of Deep-drawing of Square Blanks and the Numerical Solution of the Flange Region, Proceedings of the 11th Biennial Congress of the IDDRG, Memories Scientifiques Revue Metallurgie, No. 4 (April 1980), pp. 525-534. Kalla, U. et al., Some Aspects of HSCR Sheets For Autobodies, Proceedings of the 12th Biennial Congress of the IDDRG, S. Marghertia Lilgure, Italy, (May 24-28, 1982), pp. 345-354. Kalpakjian, S., Recent Progress in Metal Forming Tribology, Journal of Applied Metal Working, 4, No. 3, (July 1986), pp. 270-280. Kasper, A. S., How We Will Predict Sheet Metal Formability, Metal Progress, 96, (October 1969), pp. 159-164. Kasper, A. S., How Steels and Dies Interact in Forming Shapes, Metal Progress, 99, (May 1971), pp. 57-60. Kasper, A. S. and VanderVeen, P. J., A New Method of Predicting the Formability of Materials, S'E Paper No. 720019, (January 1972). Kasper, A. S., et al, Sheet Metal Forming Limits With Manufacturing Applications, Discussion to Advances In Deformation Processing, 21st Sagamore Conference, Plenum Press, New York, (1974), Unpublished. Kawai, N. and Dohda, K., Triboloby in Metal Forming Processes, Advanced Technology Of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 145-156. Keeler, S. P., Automotive Sheet Metal Formability, Research Report SG79-1, American Iron and Steel Institute, (July 1979), pp. 53-57. Keeler, S. P., Circle Grid Analysis (GCA), National Steel Booklet, Livonia, Michigan, (1986). Keeler, S. P., Statistical Deformation Control for SPQC

K_12 K_13

K_14

K_15

K_16

K_17

K_18 K_19

K_20

K_21

K_22

K_23

K_24

Monitoring of Sheet Metal Forming, S'E Paper No. 850278, (1985). Keeler, S. P., Determination of Forming Limits in Automotive Stampings, S'E Paper No. 650535, (May 17, 1965). Keeler, S. P., Evaluating the Formability of Galvanized Steels, Detroit ASM Conference on Solutions to Manufacturing Problems Using Galvanized Steels, Summary by W. J. Riffe, AISI, Washington, D.C., (1985), pp. 14-16. Keeler, S. P. and Backofen, W. A., Discussion to R. L. Whiteley, The Importance of Directionality in Drawing Quality Sheet Steel, Trans. ASM. 52, (1960), pp. 166-168. Keeler, S. P. and Backofen, W. A., Plastic Instability and Fracture in Sheets Stretched Over Rigid Punches, ASM Trans. Quarterly, 56, (March 1964), pp. 25-48. Keeler, S. P. and Dwyer, T. E., Frictional Characteristics of Galvanized Sheets Evaluated with a Draw Bead Simulator, S'E Paper No. 860433, (1986). Keeler, S. P., Evaluating the Lubricity of Press Shop Lubricants on Various Types of Galvanized Steels, AIME Mechanical Working Conference, Greentree, Pittsburgh, PA, (October 17, 1986). Keeler, S. P., Use of Grid Systems for Strain Determinations, Paper for IDDRG Working Group I, London, (June 5, 1964). Keeler, S. P., New Sheet Metal Qualities and Their Evaluation, International Congress System Sheet 75, Zurich, Switzerland, (September 16, 1975). Keeler, S. P., Ductility of Anisotropic Sheet Metal, Chapter 8, Ductility, American Society for Metals, Metals Park, Ohio, (1967), pp. 227-254. Keeler, S. P., Circular Grid System A Valuable Aid for Evaluating Sheet Metals Formability, S'E Paper No. 680092, (January 1968). Keeler, S. P., Interactions Between Surfaces and Formability, Technological Impact Of Surfaces: Relationship To Forming, Welding, And Painting, American Society for Metals, Metals Park, Ohio, (1982). Keeler, S. P., Sheet Metal Stamping Technology Need for Fundamental Understanding, Mechanics of Sheet Metal Forming Material Behavior And Deformation Analysis, General Motors Research Symposium, Plenum Press, New York, (1978), pp. 3-17. Keeler, S. P., Understanding Sheet Metal Formability, Machinery, February 68: pp. 88-95, March 68: pp. 94-101, April 68: pp. 94-103, May 68: pp. 92-99, June 68: pp. 98-104, July 68: pp. 78-83. Also Sheet Metal Industries, 48, (1971), Issues No. 5-

K_25

K_26

K_27

K_28

K_29

K_30

K_31 K_32 K_33

K_34

K_35

K_36

K_37 K_38

10. Keeler, S. P., Sheet Metal Stampings-Future Design and Control, AIME Detroit Section Spring Seminar, Detroit, (March 18, 1974). Keeler, S. P., Understanding Sheet Metal Formability Vol. II, Paper No. 35A, Metal Fabricating Institute, Rockford, IL., (1970). Keeler, S. P., et al, Effect of Test Parameters on the Olsen Cup Test, National Steel Research Report, (April 1971), Unpublished. Keeler, S. P., Forming Limit Criteria-Sheets, Advanced In Deformation Processing, 21st Sagamore Conference, Plenum Press, New York, (1974). Keeler, S. P. and Brazier, W. G., Relationship Between Laboratory Material Characterization and Press-Shop Formability, Microalloying 74 New York, N.Y., (1977), pp. 517-530. Keeler, S. P., Press Shop Applications of Forming Limit Diagrams, IDDRG Working Group III, Ann Arbor, MI, (October 17, 1976). Keeler, S. P., Unpublished data. Keeler, S. P., Proposed Revision to S'E Div. 32 Subcommittee on the Rewrite of J126, Unpublished. Keeler, S. P. and Keeler, R. N., Formability of High Strength Steels Styling and Shape Limitations, International Nickel Annual Bumper Design Seminar, Troy, Michigan, (November 1979). Keeler, S. P., Stretch Forming-An Emerging Technology for the Automotive Industry, (To be Published Journal of Material Shaping Technology, ASM). Kikuma, T. and Nakazima, K., Effect of Deforming Conditions and Mechanical Properties on the Stretch Forming Limits of Steel Sheets, Proceedings of the 6th Biennial Congress of the International Deep Drawing Research Group, (1970), Supp. to Trans. of the Iron and Steel Inst. of Japan, 11, (1971), pp. 827831. Kim, D. W. and Kim, N. S., A Study of Pure Stretch Forming, Advanced Technology Of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 753-758. Kleemola, H. J. and Pelkkikangas, M. T., Sheet Metal Industries, 54, (1977), p. 591. Kleemola, H. and Backlund, U., Damage Formation in Deep Drawing of Mild and High-strength Low Alloy Sheet Steels, Proceedings of the Working Group Meetings of the IDDRG, S.

K_39

K_40

K_41

K_42

K_43

K_44

K_45

K_46

K_47 K_48

K_49

Margherita Ligure, Italy, (May 24-28, 1982), Working Group Volume, pp. 49-60. Kleemola, H. J., Korhonen, A. S., and Rranta-Eskola, A. J., Effect of Strain Rate and Deformation Heating on the StrainHardening of Sheet Metals, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17-21, 1978), pp. 175-178. Kleemola, H. J., Kumpulainen, J. O., and Ranta-Eskola, A. J., Factors Influencing the Forming Limits of Sheet Metals, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 3 (March 1980), pp. 403-413. Kleemola, H. J. and Pelkkikangas, M. T., Effect of Predeformation and Strain Path on the Forming Limits of Steel, Copper, and Brass, IDDRG WG II, Ann Arbor, MI, (1976). Kleemola, H. J. and Pelkkikangas, M. T., Effect of Stress State and Predeformation on the Strain Hardening of Steel, Copper, and Brass, IDDRG Working Group II, (1976). Klein, A. J. and Hitchler, E. W., Hole-Enlargement Practical On-Line Formability Test, IDDRG Working Group III, Amsterdam, (1972). Kobayushi, S. and Kim, J. H., Deformation Analysis of Axisymmetric Sheet Metal Forming Processes by The Rigidplastic Finite Element Method, Mechanics of Sheet Metal Forming Material Behavior And Deformation, General Motors Research Symposium, Plenum Press, New York, (1976). Kobayushi, T. et al, Effect of Strain Ratios on the Deforming Limit of Steel Sheet and Its Application to Actual Press Forming, in Sheet Metal Forming And Formability, Proceedings of the 7th Biennial Congress of the International Deep Drawing Research Group, Amsterdam, (October 9-13, 1972), pp. 8.1 to 8.4. Kobayushi, T.. et al, Relations Between Deforming Behaviors and Geometrical Factors of Autobody Parts in Press Forming, Proceedings of the 6th Biennial Congress of the International Deep Drawing Research Group, (1970), Supp. to Trans. of the Iron and Steel Inst. of Japan, II, (1971), pp. 841-845. Kocks, U. F., Lecture on Strain Rate Hardening, Detroit Section AIME, Fall 1977. Koistinen, D. P. and Wang, N. M. editors, Mechanics Of Sheet Metal Forming-Material Behavior And Deformation Analysis, General Motors Research Symposium, Plenum Press, New York, (1978). Kojima, M. et al, Effectiveness of Flange Holding on the Die Surface with Draw Beads, Sheet Metal Formability And Energy

K_50

K_51

K_52

K_53

K_54

K_55 K_56

K_57

K_58

L_1

L_2

Conservation, Proceedings of the 9th Biennial Congress of the International Deep Drawing Research Group, Ann Arbor, MI, (October 13-14, 1976), pp. 207-219. Kokkonen, V., Modeling of Forming Processes For Tool Design and Manufacturing at Volvo,Computer Moedling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 13-20. Korhonen, A. S. and Sulonen, M., Force Requirements In Deep Drawing of Cylindrical Shells, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 4 (April 1980), pp. 515-524. Korhenen, A., Sirvio, E., and Sundquist, H., New Ways of Improving the Surface Properties of Metal Forming tools, Proceedings of the 12th Biennial Congress of the IDDRG, S. Margherita Ligure, Italy, (may 24-28, 1982), pp. 175-182. Korhonen, A. S., Effect of Forming and Testing Speed on Formability and Mechanical Properties of Sheet Metal, Proceedings of the Working Group Meetings of the IDDRG in S. Margherita Ligure, Italy, (May 1982), Working Group Volume pp. 191-196. Kozono, H., Computer-aided Profile Measurement and Analysis of Pressed Steel Sheets, Proceedings of the Working Group Meeting of the IDDRG, S. Margherita Ligure, Italy, (May 24-28, 1982), Working Group Volume pp. 73-82. Krupitzer, R. P., et al, Progress in HSLA Steels in Automotive Applications, S'E Paper No. 770162, (1977). Kubotera, H. and Ueno, Y., Influence of Strain History on Formability of Steel Sheet, NKK Overseas Technical Report, (March 1970), pp. 19-25. Kula, B. and Fahey, N. H., Effect of Specimen Geometry on Determination of Elongation in Sheet Tensile Specimens, Materials Research and Standards, 1, (1961), p. 631-636. Kumpulainen, J. O., Factors Influencing Friction and Galling Behavior of Sheet Metals, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (1984), pp. 476490. Lahoti, G. D., Nagpal, V., and Altan, T., Prediction of Limit Strains in Biaxial Stretching of Sheets Using Marciniaks Approach, Advanced Technology Of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 712-718. Lake, J. S. H., The Yoshida Test A Critical Evaluation and Correlation With Low-strain Tensile Parameters, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne,

L_3 L_4

L_5 L_6

L_7

L_8

L_9

L_10

L_11

L_12

L_13 L_14 L_15

L_16

Australia, (February 20-25, 1984), pp. 554-464. Lange, K., Handbook Of Metal Forming, McGraw-Hill, (1985). Laukonis, J. V. and Ghosh, A. K., Effects of Strain Path Changes on the Formability of Sheet Metals, Met Trans. A, 9A, (Dec. 1978), p. 1849. Lee, D. and Forth, C. M., Report No. 83 CR D223, General Electric Company, New York (1983). Lee, A. P., and Hiam, J. R., Factor Influencing the Forming Limit Curves of Sheet Used for Press Forming, Canadian Institute of Metallurgy, 12th Annual Conference of Metallurgists, Quebec, (1973). Levy, B. S., Modeling Binder Restraint Using Parametric Models Based on Mechanistic Considerations, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 177191. Levy, B. S., Empirically Derived Equations for Predicting Springback in Bending, J. of Applied Metalworking, 3, No. 2, (January 1984), pp. 135-141. Levy, B. S., Development of a Predictive Model for Drawbead Restraining Force Utilizing work of Nine and Wang, Journal of Applied Metalworking, 2, No. 3, (1982), pp. 193-199. Levy, B. S., On the Possibility of a Unique Test for Sheet Metal Formability, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), p. 592. Liebig, H. P. and Beyer, R., Evaluation of Sheet Steel Material Deformability by Tensile Test Without Strain Control, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 464-465. Liege, C. R. M., Present State of the CRM Work Related to Surface Analysis of Cold-Rolled Steel Sheets, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17-21, 1978), pp. 253-279. Littlewood, M. and Wallace, J. F., Sheet Metal Industries, 41, (1964), pp. 925-930. Liu, Y. C., On the R-value Measurements Met Trans A, 14A (June 83), pp. 1199-1205. Liu, Y. C., Springback Reduction in U-Channels Double Bend Technique, J. of Applied Metalworking, 3, No. 2, (January 1984), pp. 148-156. Logan, R. W. and Hosford, W. F., Wall Wrinkling During Stretch Draw: Simulation and Experiments, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The

M_1

M_2

M_3

M_4

M_5

M_6

M_7

M_8 M_9 M_10

M_11 M_12

M_13

Metallurgical Society of AIME, Warrendale, PA (1986), pp. 225242. MacPhee, J., An Engineering Analysis of the Redrawing Process (Part 2), Sheet Metal Industries, 54, (Jan. 1977), pp. 75-78. Makimattila, S. and Ranta-Eskola, A., Behavior of Galvanized Coatings During Forming, Proceedings of the 13th Biennial Congress of the International Deep Drawing Research Group, Melbourne, Australia, (February 1984), pp. 293-304. Makimattila, S. and Ristolainen, E., Effect of Surface Chemical Condition on the Frictional Properties of Hot-Dip Galvanized Steel Sheets, Proceedigns of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 261-270. Makinouchi, A., Elastic-Plastic Stress Analysis of U-Bend Process of Sheet Metal, Advanced Technology Of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 672-677. Makinouchi, A., Elastic-Plastic Stress Analysis of Bending and Hemming of Sheet Metal, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 161-176. Marciniak, Z., Sheet Metal Forming LImits, Mechanics Of Sheet Metal Forming-Material Behavior And Deformation Analysis,General Motors Research Symposium, Plenum Press, New York, (1978), pp. 215-233. Marciniak, Z. et al, Influence of the Plastic Properties of a Material on the Forming Limit Diagram for Sheet Metal in Tension, Int. J. Mech. Sci., 15, (1973), pp. 789-805. Marciniak, Z., Aspects Of Material Formability, McMaster University, Hamilton, Ontario, (1974). Marciniak, S., Kuczynski, K., and Pokora, T., J. Mech. Sci., 15, (1973), p. 789. Marciniak, Z., Assessment of Material Formability, Advanced Technology Of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 685694. Marciniak, Z. and Kuczynski, K., J. Mech. Sci., 9, (1967), p. 609. Matsudo, K., Osawa, K., and Yoshida, M., Prediction of Stretch Flagability of Punched Surface of Steel Sheets, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17-21, 1978), pp. 325. Matsui, M. and Iwat, N. Buckling and Its Growth in Biaxial Diagonal Stretch on Square Plates of Steel Sheet, Proceedings

M_14

M_15 M_16

M_17

M_18

M_19

M_20 M_21

M_22 M_23

M_24

M_26 :

M_27 : M_28 :

M_29 :

of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 521-531. Matsuoka, T., and Yamamori, K., Metallurgical Aspects in Cold Rolled High Strength Steel Sheets, ASM Materials Engineering Congress, Chicago, Illinois, (October 1973). Matsuoka, T. and Sudo, C., The Sumitomo Search No. 1, (1969), p. 71. Melander, A., The Forming Limit Diagram of High Strength Materials, Proceedings of the Working Group Meetings of the IDDRG, S. Margherita Ligure, Italy, (May 24-28, 1982), Working Group Volume, pp. 17-30. Melin, L E. and Nilsson, K. I., Frictional Forces In Pressforming of Laminated Steel, Proceedings of the 12th Biennial Congress of the IDDRG, S. Margherita Ligure, Italy, (May 24-28, 1982), pp. 287-298. Mellor, P. B., Deep-Drawing and Stretch Forming, Developments In The Drawing Of Metals, Proceedings of the International Conference, the Metals Society, London, (1983), pp. 76-86. Mellor, P. B. and Parmar, A., Plasticity Analysis of Sheet Metal Forming, Mechanics Of Sheet Metal Forming Material Behavior And Deformation Analysis, General Motors Research Symposium, Plenum Press, New York, (1978), pp. 53-74. Mellor, P. A., Inc. Metals review, No. 1, (1981), pp. 1-20. Menzer, W., Chemical Surface Treatment and Corrosion Protection in the Automotive Industry, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 63-114. Meuleman, D. J. and Brazier, W. G., Oral discussion at the 1986 S'E Congress, Detroit, (February 1986). Meuleman, D. J., Siles, J. L., and Zoldak, J. J., The Limiting Dome Height for Assessing the Formability of Sheet Steel, S'E Paper No. 850005, (1985). Meuleman, D. J. and Zoldak, J. J., The Interactions of Coated Steels, Die Materials, and Forming Lubricants, S'E Paper No. 860432, (1986). Meuleman, D. J., Denner S. G., and Cheng, F. L., The Effect of Zinc Coatings on the Formability of Automotive Sheet Steels, S'E Paper 840370, (1984). Meuleman, D. J., Unpublished Research. Meyer, L. et al, Possibilities of Improving the Surface Cleanliness of Sheet Steel. Results and Critical Remarks, Proceedings of the Working Group Meetings of the IDDRG in S. Margherita Ligure, Italy, (May, 1982), pp. 2-5-213. Mital, N. K., Prediction of Binder Wrap in Sheet Metal

M_30 :

M_31 :

M_32 :

M_33 :

M_34 :

M_35 :

M_36 :

M_37 : M_38 : M_39 : N_1 :

N_2 :

N_3 :

Stampings Using Finite Element Method, Proceedings Of The 1983 International Conference On Computers In Engineering ASME, Vol. 3, (1983), pp. 71-77. Miyahara, M., Iwaya, J., Shirasawa, H., and Korida, K., Studies on Formability of 1,000 MPa Cold Rolled Steel Sheet, pp. 402403. Miyauchi, K., Stress-Strain Relationship in Simple Shear of InPlane Deformation for Various Steel Sheets, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 436-443. Miyauchi, K., Torsional Behavior of Sheet Metals, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 400-401. Miyauchi, K., Furubayashi, T., Herai, T., and Sudo, Co. in Sheet Metal Forming And Formability, Proceedings of the 10th Biennial Congress of the IDDRG, Portcullis Press, Redhill, Surrey, England, (1978), pp. 287-301. Mohammed, Ali, A. and Mellor, P. B., Constitutive Relations and Forming Limit Curves, Advanced Technology of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 701-706. Monfort, G. and Bragard, A., A Simple Model of Shape Errors in Forming and Its Application to the Reduction of Springback, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 273287. Mould, P. R. and Johnson, T. E., Rapid Assessment of Drawability of Cold-rolled Low-carbon Steel Sheets, Sheet Metal Industries, 50, (June 1973), pp. 328-348. MTIA Production Technology Center, A Training Course On Sheet Metal Forming 8th Edition, Australia, (September 1977). uchenborn, W. and Sonne, H., Archieve fur das Eisenhuttenwegen, 46, (1946), p. 597. Muchenborn, W. and Sonne, H. M., Material Properties Controlling the Strain Distribution of Sheet Metal. Murakawa, M., Burr-Free Shearing, Advanced Technology Of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 805-814. Nagpal, V., Billhardt, C. G. F., Raghupathi, P. S., and Subramanian, T. L., Development of a Process Model for Press Brake Bending, Process Modeling Fundamentals And Applications To Metals, American Society for Metals, Metals Park, Ohio, (1980), pp. 287-301. Nakagawa, K. and Abe, H., Press Formability of High Strength

N_4 :

N_5 :

N_6 :

N_7 :

N_8 :

N_9 :

N_10 :

N_11 :

N_12 :

N_13 : N_14 :

N_15 :

Cold Rolled Steel Sheets, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 4 (April 1980), pp. 475-484. Nakawaga, T. and Yoshida, K., Improvement in the Deformability of Sheared Edge of Steel Sheets, Proceedings of the 6th Biennial Congress of the International Deep Drawing Research Group, (1970), Supp. to Trans. of the Iron and Steel Inst. of Japan, II, (1971), pp. 813-818. Nakamachi, E. and Sowerby, R., Numerical Analysis of Bulging and Punch Stretching of Circular Disks Based on Kirchhoff Plate Theory, Advanced Technology of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 666-671. Nakazima, K. et al, Effects of Mechanical Properties on the Press Forming Severity of Steel Sheets, IDDRG Working Group II, Amsterdam, (1972). Nakazima, K., Kikuma, T., and Hasuka, K., Study on the Formability of Steel Sheets, Yawata Technical Report No. 264, Sept. 1968. National Steel Corporation, Relative Formability of High Strength Steels-A Case Study with Bumpers, Automotive Tehcnical Bulletin No. FO-4G9, (1969). Naziri, H. and Pearce, R., The Effect of Plastic Anisotropy on Flange-Wrinkling Behavior During Sheet Metal Forming, Int. J. Mech. Sci., 10, (1968), pp. 681-694. Neda et al, A Study of Springback in the Stretch Bending of Channels, J. of Mechanical Working Technology, 5, (1981), pp. 163-179. Needleman, A. and Rice, J. R., Limits to Ductility Set By Plastic Flow Localization, Mechanics Of Sheet Metal FormingMaterial Behavior And Deformation Analysis, General Motors Research Symposium, Plenum Press, New York, (1978), pp. 237-265. Neubauer, A. and Rasehorn, H. J., New Technological Research on Testing Methods and Cold-Bending of Coated Steel Sheets, Proceedings of the 14th Biennial Congress of the International Deep Drawing Research Group, Koln, Germany, (April 1986), pp. 295-304. Newby, J. R., A Practical Look at Biaxial Stretch Cupping Tests, Sheet Metal Industries, 54, (March 1977), pp. 240-252. Nihill, T. J. and Thorpe, W. R., Case Studies Using Grid Square Strain Analysis, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 23-33. Nine, H. D., Drawbead Forces in Sheet Metal Forming,

O_1 :

P_1 :

P_2 :

P_3 :

Q_1 :

R_1 : R_2 :

R_3 :

R_4 :

R_5 : R_6 :

R_7 : R_8 :

Mechanics Of Sheet Metal Forming-Material Behavior And Deformation Analysis, General Motors Symposium, Plenum Press, New York, (1978), pp. 179-207. Osakada, K. and Murayama, F., A Method for Evaluating Tool Materials Against Seizure in Metal Forming, Advanced Technology of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 168173. Painter, M. J. and Pearce, R., Instability and Fracture in Sheet Metal, in Sheet Metal Forming And Formability, Proceedings of 7th Biennial Congress of the International Deep Drawing Research Group, Amsterdam, (October 1972), pp. 1.1 to 1.4. Pearce, R., The Effect of Varying Strain-Ratio on the Hydraulic Bulging Behavior of Sheet Metals, Colloquium on The Influence Of Material Properties On The Forming Of NonFerrous Metal Sheet, BDDRG, (April 11, 1967), pp. 1-29. Pirvu, G., The Deformation and Fracture Mechanism of Dual Phase Steel Sheets Under Biaxial Stress Combinations, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 254-262. Queener, C. A. and DeAngelis, R. J., Elastic Springback and Residual Stresses in Sheet Metal Formed by Bending, ASM Transactions, 61 (1968), pp. 757-768. Rabinowicz, E., Friction And Wear Of Materials, Wiley, New York, 1965. Ranta-Eskola, A., Effect of Loading Path on the Formability of Sheet Metals, Proceedings on the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 4 (April 1980), pp. 543-552. Rao, U. S. and Chaturvedi, R. C., Effect of Sheet Metal thickness on Forming Limits, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 85-94. Rao, U. S. and Chaturvedi, R. C., Effect of Sheet Metal Thickness on Forming Limits of Strain Rate Sensitive Material, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 420-421. Rashid, M. S., GM980X-Potential Applications and Review, S'E Paper No. 770211, (1977). Rasmussen, G. K., Formability Limit Curves and Their Relationship to Lubrication, Paper present to ADDRG, Wilmington, Del., (1973). Rasmussen, S. N., CIRP Annals, 50, No. 1, (1981), p. 179. Rasmussen, S. N., A Shear Test for Sheet Metal Bendability, Proceedings of the 13th Biennial Congress of the IDDRG,

R_9 :

R_10 :

R_11 :

R_12 :

R_13 :

R_14 :

R_15 :

R_16 : R_17 : R_18 :

R_19 :

S_1 : S_2 :

Melbourne, Australia, (February 20-25, 1984), pp. 415-526. Rasmussen, S. N., Assessing the Influence of Strain Path on Sheet Metal Forming Limits, Proceedings of the Working Group Meetings of the IDDRG, S. Margherita Ligure, Italy, (May 24-28, 1982), Working Group Volume, pp. 83-94. Ratke, L. and Welch, P. I., The Differential Work Hardening Coefficient and Classical Work Hardening Laws, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 427-435. Rault, D. and Entringer, M., Anti-Galling Roughness Profile permitting Reduction of the Blankholder Pressure and the Required Amount of Lubricant During Forming of Sheets, Sheet Metal Forming And Energy Conservation, Proceedings of the 9th Biennial Congress of the IDDRG, Ann Arbor, MI, USA, (1976), pp. 97-114. Rault, D. and Entringer, M., Relative Sheet/Tool Velocity: Its Influence on Friction and Forming Behavior of Coated Steels, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 456-457. Reissner, J. and Schmid, W., The Importance of Materials Data in the Simulation f Sheet Forming Process, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 329-340. Rhodes, A., Computer Aided Design and Manufacture of Rolls for Cold Roll Forming Machines, SME Conference on Technology of Roll Forming, Toronto, Ontario, Sept. 25-27, 1984. Richmond, O, Devenpeck, M. L., and Marison, H. L., A Tapered Tension Specimen for Measuring Sheet Stretchability, Novel Tehcniques In Metal Deformation, R. H. Wagoner ed. TMS of AIME, Warrendale, PA, (1983), pp. 89-98. Riesz, C. H. in Metal Deformation Processes: Friction And Lubrication, Schey, J. A. (ed), Dekker, New York (1970). Roper, R. E., The Wallace Expanding Process, S'E Paper No. 660067, (January 1966). Rose, F. A., Lders Band patterns and Their Effect on the Properties and Performance of Metallic Coated Steel Strip, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 339-350. Rydell, F., Fluid Forming of Sheet Metal in Production, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 95-105. Sachs, G., Metallwirtschaft, 9, (1930), pp. 213-218. Sachs, G., New Researches on the Drawing of Cylindrical Shells, Proc. Inst. Automobile Engineers, London, 9, (1934-35),

S_3 : S_4 :

S_5 :

S_6 :

S_7 :

S_8 :

S_9 :

S_10 :

S_11 :

S_12 : S_13 : S_14 :

S_15 :

S_16 :

pp. 588-600. Sachs, G., Principles And Methods Of Sheet-Metal Fabricating, Reinhold Publishing Corp., N.Y., (1966). Saka, K., Painter, M. J., and Pearce, R., The Uniaxial StrainHardening Behavior of Sheet Metal From Zero Strain to Failure, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17-21, 1978) pp. 167-174. ang, H. and Mishikawa, Y., A Plane Strain Tensile Apparatus, Novel Techniques in Metal Deformation, R. H. Wagoner ed., TMS of AIME, Warrendale, PA, (1983), pp. 3-14. Sardeshpande, M. R., Applications of Cold Punch Process and Edge Stretch Analysis to Eliminate Failure Conditions in Forming of Sheet Steels," S'E Paper No. 860128, (1986). Sata, S, Shiokawa, M., Furubayashi, T., and Yamazaki, K., Required Properties of High Strength Steel Sheets for Autobody Parts. Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 3 (March 1980), pp. 237-246. Sawada, T., Measurements of the Dynamic Stress Time Curves with Good Accuracy, Proceedings of the Working Group Meetings in S. Margherita, Italy, (May 1982), Working Group Volume, pp. 159-170. Saxena, A. and Chatfield, D. A., High Strain Rate Behavior of Some Hot and Cold Rolled Low Carbon Steels, S'E Paper No. 760309, (1976). Schedin, E., Karlsson, S., and Melander, A., Damage Development in a Hot-Dip Galvanized Coating During Forming, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 460-461. Schey, J. A., Tribology In Metalworking: Friction, Lubrication, And Wear, American Society for Metals, Metals Park, Ohio, (1983). Schey, J. A. (ed), Metal Deformation Processes: Friction And Lubrication, Dekker, New York (1970). Scroggins, R., Problems in Sheet Metal Forming, Paper No. F1200, Fabrication Machinery Association, Rockford, IL, (1973). Semiatin, S. L. and Jonas, J. J., Formability And Workability Of Metals: Plastic Instability And Flow Localization, ASM, Metals Park, Ohio, (1984). Segala, A. and Pearce, R. A Study on Compressive Instability Using the Yoshida Test, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 476-477. Shaffer, S., The Relevance of Inclusions on Formability in Punch-Stretching of Low-Carbon, AK, DQ Steels, M. S. Thesis

S_17 :

S_18 :

S_19 :

S_20 :

S_21

S_22

S_23 S_24

S_25

S_26 S_27

S_28

and Report LBL-21742, Lawrence Berkeley Laboratory, University of California, (May 1986). Shimomura, T. and Yoshida, M., Study of Surface Deflection in Door Outer Panel, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 46-53. Shimomura, T., Yoshida, M., Sakoh, M., and Matsudo, K., Formability of High Strength Cold Rolled Steel Sheets With 400 Mpa Tensile Strength, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 4 (April 1980), pp. 465-474. Shinozaki, M., Matsumoto, Y., Tsunoyama, K., and Irie, T., Formability of Steel-Plastic-Steel Laminate, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21-23, 1986), pp. 159-168. Shiokawa, M., Tzkizawa, H., Ujihara, S., and Hayashi, Y., A New Method for Evaluating the Powdering Phenomenon of Precoated Sheet Metals, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20-25, 1984), pp. 305-316. Shiokawa, M., Takizawa, H., Ujihara, S., Usuda, M., and Ohwue, T., Predictive Evaluation of Draw_in, Elongation, and Breakage During Forming Processes, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 243_257. Shutack, D. S. and Hultgren, F. A., The Mechanics of Die Bending Part I, Republic Steel Research Lab Report, (May 13, 1969), publication unknown. Sidey, M. P., Determination of r Values By A Weighing Technique, British Steel Corp. Report SM/TN/A/1145, (1974). Siekirk, J. F., Process Variable Effects on Sheet Metal Quality, Journal of Applied Metalworking, 4, No. 3, (July 1986), pp 252_269. Society of Automotive Engineers, Properties of Low Carbon Steel Sheets and Strips and Their Relationship to Formability, S'E J877, S'E Handbook, S'E, New York. Society of Automotive Engineers, Handbook Supplement 30 Iron And Steel, Printed by AISI, Washington, D.C., (1977). Society of Automotive Engineers, Selecting and Specifying Hot and Cold Rolled Steel Sheet and Strip S'E J126 June 81, S'E Handbook, S'E, New York. Society of Automotive Engineers, Method of Determining Plastic Deformation in Sheet Steel Stampings S'E J863 Dec. 81, S'E Handbook, S'E, New York.

S_29

S_30

S_31

S_32 S_33 S_34 S_35 S_36 S_37 S_38

S_39 S_40 S_41 S_42

S_43

Society of Manufacturing Engineers, Tool And Manufacturing Engineers Handbook, Vol. 2 (Forming), Chapter 8 (Roll Forming), SME, Dearborn, MI, 1984, pp. 8_1 to 8_33. Society of Manufacturing Engineers, Tool And Manufacturing Engineers Handbook, Vol 2 (Forming), Chapter 1 (Sheet Metal Forming), pp. 1_1 to 1_33. Souza_Nobrega, C., Fidelis DaSilva, B., Ferran, G., Jalinier, J. M., and Baudelet, B., Effect of the Inclusions Content on Damage Generation During Sheet Metal Forming of Low Carbon Steel, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 3 (March 1980), pp. 293_302 Sowerby, R., Chu, E., and Duncan. J. L., J. Strain Analysis, 17, (1982), pp. 95_101. Sowerby, R., Duncan, J. L., and Sklad, M. P., Sheet Metal Industries, (May 83), pp. 290_293. Sowerby, R. and Duncan, J. L., Failure in Sheet Metal in Biaxial tension, Int. J. Mech. Sci., 13, (1971), pp. 217_229. Sowerby, R. and Sareen, B. K., Formability and Fracture of Some Higher Strength Aluminum Alloys, Publication unknown. Stevenson, R., Journal of Applied Metal Working, 3, (1984), p. 272. Stevenson, R., Formability of Galvanized Steels Revistied, S'E Paper No. 850276, (1985). Stine, P. A., Seward, R. E., Beyerle, M. T., and Luken, P. C., C'E Sheet Metal Formability Model Predictive Capability Improved with Experimentally Derived Input Data, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 107_120. Stine, P., Basic Design Guidelines for Sheet Metal Parts, Unpublished. Stine, P., Unpublished Test Results. Stine, P. A., Sheet Metal Formability Analysis, Proc. of Porcelain Enamel Institute, 45, (1973), pp. 19_36. Stoughton, T. B., Finite Element Modeling of 1008 AK Sheet Steel Stretched Over a Rectangular Punch with Bending Effects, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp.143_159. Sudoo, C., Hayashi, Y., and Nishihara, M., Behavior of Coated File in Press Forming of Surface Treated Steel Sheet, Memoires Scientifiques Revue de Metallurgie, No. 3, (March 1980), pp.

S_44 S_45

S_46

T_1 T_2

T_3

T_4

T_5

T_6

T_7

T_8 T_9

353_362. Swift, H. W., Plastic Instability Under Plane Stress, J. Mech. Phy. Salids, 1, (1952), p.1. Swift, H. W. and Chung, S. Y., Cup Drawing From a Flat Blank Part I, Experimental Investigatins, Part II, Analytical Investigations, Proc. Inst. Mech. Eng., 165, (1951), p. 199. Szacinski, A M. and Thomson, P. F., The Effect of Mechanical Properties on the Wrinkling Behavior of Sheet Materials in the Yoshida test and in a Cone Forming Test, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20_25, 1984), pp. 532_543. Takahashi, A., Okamoto, I., and Mori, T., Proc. CAMP 83, (1983), p. 623. Takahashi, A., Okamoto, I., Hiramatsu, T., and Yamada, N., Evaluation Methods of Press Forming Severity In CAD Applications, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 37_50. Takahina, K., Herai, T., Komorida, H., and Horita, T., Relation Between the Manufacturing Conditions and the Average Strain According to the Scribed Circle Tests in Steel Sheet, La Metallurgia Italiana, 60, No. 8, (1968), pp. 757_765. Tang, S. C., Verification and Application of a Binder Wrap Analysis, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 193_208. Thomson, P. F., A Theoretical Approach to the Development of Forming Limit Criteria in Plastic Working Processes, Advanced Technology Of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 719_722. Thomson, P. F., The Roughening of Free Surfaces Under Deformation, Proceedings of the Working Group Meetings in S. Margherita, Italy, (May 1982), Working Group Volume, pp. 123_128. Thomson, P. F. and Nayak, P. U., The Development of Thickness Non_uniformities Leading to Failure in Sheet Metals, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifques Revue Metallurgie, No. 3 (march 1980), pp. 303_312. Thomson, T. R., et al., The Formability of Sheet Steels, BHP Tech. Bull. 18, No. 2, (Nov. 1974), pp. 15_19. Thomson, T. R., Influence of Material Properties in the

T_10

T_11

T_12

T_13

T_14

T_15 T_16

T_17

U_1

V_1 V_2 V_3

V_4

Forming of Square Sheels, Journal of the Australian Institute For Metals, 20, No. 2, (1975), pp. 106_113. Thomson, T. R. and Hobbs, R. M., Dual Phase Steels Production and Formability, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 4 (April 1980), pp. 455_464. Thorpe, W. R. and Nihill, T. J., Sheet Metal Strain Analysis System Based on a Square Reference Grid, Journal of Mechanical Working Technology, (1984). Thorpe, W. R. and Nihill, T. J., Substitution of a Low Carbon Steel With a Re_phosphorized Variety for an Automotive Panel: The Application of Grid Squares Strain Analysis, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20_25, 1984), pp. 34_45. Tozawa, Y., Plastic Deformation Behavior Under Conditions of Combined Stress, Mechanics Of Sheet Metal Forming Material Behavior And Deformation, General Motors Research Symposium, Plenum Press, New York, (1978), pp. 81_109. Tozawa, Y.et al., Yield Loci for Prestrained Steel Sheets, Proceedings of the 6th Biennial Congress of the International Deep Drawing Research Group, (1970), Supp. to Trans. of the Iron and Steel Inst. of Japan, II, (1971), pp. 936_940. Truszkowski, W., Arch. Hutn., 4, (1959), p. 283. Truszkowski, W., On the Quantitative Evaluation of Plastic Anisotropy in Sheet Metals, Proceedings of the 8th Biennial Congress of the International Deep Drawing Research Group, Gothenburg, Sweden, (1974), pp. 48_61. Truszkoswki, W., On Physical Meaning of the Stress_Strain Relationship Parameters in High Strength Polycrystalline Metals, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 3 (March 1980), pp. 193_202. Umehara, Y., Analysis of Shape Fixability in High Tensile Strength Steel Sheets in Press Forming, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 3 (March 1980) and No. 4 (April 1980), pp. 247_256. Vaccari, J. A., Forming and Welding High_Strength Sheet Steels, American Machinist, (May 1984), pp. 110_124. VanderVeen, P. J. and Kasper, A. S., Chrysler Shape Analysis, ADDRG Cooperative Program, (July 30, 1971). Vecchio, M. T., Design Analysis and Behavior of a Variety of As_Formed Mild and High Strength Sheet Materials in Large Deflection Bending, S'E Paper 830398, (1983). Veerman, C. C., Hartman, L., Peels, J. J., and Neve, R. F.,

V_5

V_6

W_1

W_2

W_3

W_4

W_5

W_6

W_7

W_8

W_9

W_10

Determination of Appearing Strains and Admissible Strains in Cold_Reduced Sheets, Sheet Metal Industries, 48, No. 9, (1971), pp. 678_680, 692_694. Venter, R. D. and DeMalherbe, M C., Theoretical Estimate of the Keeler_Goodwin Formability Curve, Sheet Metal Industries, 48, No. 9, (1971), pp. 656_658. VonStebut, J. and Moulene, B., The Choice of Meaningful Parameters Obtained From Surface Profile Analysis in Order to Characterize the Deep Drawing Capacity of a Metal Sheet, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20_25, 1984), pp. 504_514. Wagoner, R. H., Constitutive Equations for Sheet Forming Analysis, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application Wang and Tang, ed., The Metallurgical Society of AIME, Warrendale, PA (1986), pp. 77_92. Wagoner, R. H., Development of Constitutive Equations for Sheet Forming Analysis, Advanced Technology Of Plasticity, Proceedings of the First International Conference on Technology of Plasticity, Tokyo, (1984), pp. 707_711. Wagoner, R. H., Editor, Novel Techniques In Metal Deformation Testing, Conference Proceedings, TMS of AIME, Warrendale, PA, (1983). Wang, N. M., A Rigid_Plastic Rate_Sensitive Finite Element Method for Modeling Sheet Metal Forming Processes, in Numerical Analysis Of Forming Processes, J. F. Prittman et al eds., John Wiley and Sons, Ltd., England, (1984). Wang, N. M., A Mathematical Model of Drawbead Forces in Sheet Metal Forming, Journal of Applied Metalworking, 2, No. 3., (1982), pp. 193_200. Wang, N. M. and Wenner, M. L., Elastic_Viscoplastic Analyses of Simple Stretch Forming Problems, Mechanics Of Sheet Metal Forming_Material Behavior And Deformation Analysis, General Motors Research Symposium, Plenum Press, New York, (1978), pp. 367_398. Wang, N. M and Tang, S., Editors, Computer Modeling Of Sheet Metal Forming Processes; Theory, Verification, And Application TMS of AIME, Warrendale, PA, (1986). Wang, N. M. and Wenner, M. L., An Analytical and Experimental Study of Stretch Flanging, Int. J. Mech. Sci., 16, (1974), pp. 135_143. Wang, N. M., Predicting the Effect of Die Gap on Flange Springback, Proceedings of the 13th Biennial Congress, Melbourne, Australia, (February, 1984), pp. 133_147. Wang, N. M., and Gordon, W. J., On the Stretching of a

W_11

W_12

W_13

W_14

W_15

W_16

W_17

W_18 W_19

W_20 W_21

W_22

W_23

Circular Sheet by a Hemishperical Punch, Report No. GMR_744, General Motors Research Laboratories, Warren, MI, March 15, 1968. Weidemann, C., Deep Drawing of Coil Coated Steel Sheet, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Matallurgie, No. 3 (March 1980), pp. 343_352. Weidemann, C., Influence of Steel Sheet Topography on the Deep_Drawing Process, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17_21, 1978), pp. 245_252. Weidemann, C., The Blankholding Action of Draw Beads, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17_21, 1978), pp. 79_68. Weigel, H. U. et al, Correlation Between the Hole Enlargement Test and Conventional Mechanical_Technological Tests, IDDRG Working Group II. Welch, P. I. and Bunge, H. J., The Influence of Grain Size on the Instantaneous Anisotropy Parameter in a Low Carbon Steel, Proceedings of the 13th Biennial Congress of the IDDRG, Melbourne, Australia, (February 20_25, 1984), pp. 436_443. Wenner, M. L., On Work Hardening and Springback in Plane Strain Draw Forming, J. Applied Metalworking, 2.4 (1983) pp. 277_288. Whiteley, R. L., Anisotropy in Relation to Sheet Processing, Fundamental Of Deformation Processing, Proceedings of the 9th Sagamore Conference, Syracuse University Press, (1964), pp. 183_198. Whiteley, R. L., The Importance of Directionality in Drawing Quality Sheet Steel, Trans. ASM, 52, (1960), pp. 154_168. Whiteley, R. L. and Wise, D. E., Relationship Among Texture, Hot Mill Practice and the Deep Drawability of Sheet Steel, AIME Flat Rolled Products III, Metallurgical Society Conferences, Vol. 16, Chicago, (January 17, 1962), pp. 47_64. Wilson, D. V. et al., A Single_Blank Test for Drawability, Sheet Metal Industries, 43, (1966), pp. 465_476. Wilson, D. V., Controlled Directionality of Mechanical Properties in Sheet Metals, Metallurgical Reviews, Review 139, pp. 175_188. Wilson, W. R. D., Friction and Lubrication in Sheet Metal Forming, Mechanics of Sheet Metal Forming_Materials Behavior And Deformation Analysis, General Motors Research Symposium, Plenum Press, New York, (1978), pp. 157_174. Wollrab, P. M. and Streidl, M., Application of High_Strength Steel Sheets in Autobody Pressing Chances and Limitations,

W_24

W_25

W_26

W_27 W_28 W_29 Y_1

Y_2

Y_3 Y_4

Y_5 Y_6

Y_7

Y_8

Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21_23, 1986), pp. 149_158. Wollrab, P. M. and Drecker, H., Application of High Strength Steel Sheets for Outer Panels of Autobodies, Proceedings of the 12th Biennial Congress of the IDDRG, S. Margherita Ligure, Italy, (May 24_28, 1982), pp. 287_298. Wong, N. K., Arkun, B. H., Roberts, W. T., and Wilson, D. V., The Formability of Steel Sheets at Moderately Elevated Temperatures, Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 3 (March 1980), pp. 413_422. Woo, D. M., Analysis of Axisymmetric Forming of Sheet Metal and Hydrostatic Bulging Process, Int. J. Mech. Sci., (6), (1964), pp. 303_317. Woo, D. M., On the Complete Solution of the Deep Drawing Problem, Int. J. Mech. Sci., 10, (1968), p. 83. Wojtowicz, W. J., Lubrication Eng., 11, (1955), pp. 174_177. Woska, R., Bander Bleche Rohre, 20, (1979), pp. 562_564. Yamamoto, T., Takano, T., Oikawa, H. Abe, K., and Ohshima, S., Problems of High Strength Steel Sheets in Autobody Production Proceedings of the 11th Biennial Congress of the IDDRG, Memoires Scientifiques Revue Metallurgie, No. 3 (March 1980), pp. 227_236. Yamasaki, H., Nishiyama, T., and Tamura, K., Computer_Aided Evaluation Method for Sheet Metal Forming in Car Body, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21_23, 1986), pp. 373_382. Yang, C. T., Calculating Minimum Bend Radii From Ductility Ratings, Metals Progess, 98, (November 1970), pp. 107_110. Yellup, J. M., Modeling of Sheet Metal Flow Through a Drawbead, Proceedings of the 13th Biennial Congress, Melbourne, Australia, (February, 1984), pp. 166_177. Yoshida, K., ASTM Symposium, Formability 2000 AD, Chicago, (1980). Yoshida, F., Uniform Stretch Bending of Clad Sheet Metals Under Plane Strain Condition, Proceedings of the 14th Biennial Congress of the IDDRG, Koln, Germany, (April 21_23, 1986), pp. 426_428. Yoshida, K. and Miyauchi, K., Experimental Studies of Material Behavior As Related to Sheet Metal Forming, Mechanics Of Sheet Metal Forming Material Behavior And Deformation Analysis, General Motors Research Symposium. Plenum Press, New York (1978), pp. 19_49. Yoshida, K., Hayashi, H., Miyauchi, K., Hirata, M., Hira, T., and Ujihara, S., International Symposium on New Aspects of Sheet

Y_9

Y_10

Y_11

Y_12

Y_13

Y_14

Z_1

Metal Forming, Tokyo, (1981). Yoshida, K., Hayashi, H., Usuda, M., Amaike, T., and Komorida, H., Effects of Material Properties on Surface Deflection Behavior Due to Elastic Recovery in Press Forming, Advanced Technology Of Plasticity, Proceedings of the First International Conference, The Metals Society, London, (1983), pp. 747_752. Yoshida, K. et al, Significance of the X Value and the Workhardening Exponent (n) Under Equibiaxial Tension in the Assessment of Sheet Metal Formability, Sheet Metal Industries, 48, (October 1971), pp. 772_783. Yoshida, K., Miyauchi, K., Hayashi, H. and Kuriyama, S., Flange Behavior Analysis in Sheet Metal Forming of Complex Geometry, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17_21, 1978), pp. 1_28. Yoshida, K., Stato, T., Hayashi, H., Iwaki, M., and Ike, H., The Effects of Ion Implantation into Mild Steel Sheets, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17_21, 1978), pp. 279_286. Yuen, W. Y. D., CAD/CAM/C'E for Roll Forming, Proceedings of the 13th Biennial Congress, Melbourne, Australia, (February 1984), pp. 118_132. Yuton, Y., Nomura, S., and Kokubo, I., Press Formability of Damping Sheets, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17_21, 1978), 113_120. Zieger, M. and Beck_Schoenenberger, A., Relation Between Surface Morphology Characteristics and Gripping, Proceedings of the 12th Biennial Congress of the IDDRG, S. Margherita Ligure, Italy, (May 24_18, 1982), pp. 189_200.
Ziemloewocz, O. C. and Onate, E., Plastic Flow of Sxisymmetric Sheets and its Application to Sheet Metal Forming Processes, Proceedings of the 10th Biennial Congress of the IDDRG, Warwick, England (April 17_21, 1978), pp. 189_200.

Z_2

Sheet Metal Forming Definitions


This glossary, compiled by members of the North American Deep Drawing Research Group, presents terminology associated with the press forming of sheet steel and other metals. The definitions represent the consensus of a range of opinions concerning terms which previously contained much ambiguity and, more often than not, subjective attachment to either a particular metalworking process or application. Using the detailed experience of press shop workers, suppliers, and researchers in the ferrous and nonferrous industries, the NADDRG Nomenclature Subcommittee collated definitions to reflects todays usage. The glossary should prove to be a worthwhile reference resource. The North American Deep Drawing Research Group is an affiliate of ASM International, Metals Park, Ohio 44073. (A-6). Aging: A change in material property or properties with time. (See Quench Aging and Strain Aging)

Anisotropy: Variations in one or more physical or mechanical properties with direction. (See Normal Anisotropy, Planar Anisotropy, and Plastic Strain Ratio) Bend (Longitudinal): A forming operation in which the axis is perpendicular to the rolling direction of the sheet. Bendability: The ability of a material to be bent around a specified radius without fracture. Bending Under Tension: A forming operation in which a sheet is bent with the simultaneous application of a tensile stress perpendicular to the bend axis. Bend Radius: The radius measured on the inside of a bend which corresponds to the curvature of a bent specimen or the bent area in a formed part. Bend Test: Evaluation of a sheet metals response to a bending operation, such as around a fixed radius tool. Biaxial Stretchability: The ability of sheet material to undergo deformation by loading in tension in two directions in the plane of the sheet. Blank: The piece of sheet metal, produced in cutting dies, that is to be subjected to further press operations. A blank may have a specific shape developed to facilitate forming or to eliminate a trimming operation subsequent to forming. (See Blank Development) Blank Development: The process of determining the optimum size and shape of a blank for a specific part. Blank Gridding: Imprinting a metal blank with a pattern on the sheet surface, such pattern to be used for subsequent strain measurement. (See also Gridding) Blank Holder: That part of a forming die which holds the blank by pressure against a mating surface of the die to control metal flow and prevent wrinkling. The blank holder is sometimes referred to as Holddown or binder area. Pressure applied by mechanical means, springs, air, or fluid cushions. < b> The pressure exerted by the blank holder against the blank. This is normally adjustable to control metal flow during the drawing. Blanking: The act of cutting a blank. Brake Forming: A forming process in which the principal mode of deformation is bending. The equipment used for this operation is commonly referred to as a press brake.

Brake Press: A form of open frame, single action press comparatively wide between the housings with bed designed for holding long narrow forming edges or dies. It is used for bending and forming strips and plates. Breaking Load: The load acting on a specimen at the moment of fracture. Breaking Stress: The stress at which a material separates into two pieces, measured by a suitable device and usually reported in pounds per square inch of minimum cross section, or other equivalent. (See also: Fracture Stress) Brittleness: A tendency to fracture without appreciable plastic deformation. Brittle Fracture: A fracture which occurs without appreciable plastic deformation. Bulge Test: A test wherein the blank is clamped securely around the periphery and by means of hydrostatic pressure the blank is expanded. The blank is usually gridded so that the resulting strains can be measured. This test is usually performed on large blanks of 8-12 inches diameter. Clamping Pressure: Pressure applied to a limited area of the sheet surface, usually at the periphery, to control or limit metal flow during forming. Coil Weld: A welded joint connecting the ends of two coils to form a continuous strip. Coining: A process of cold forming metals, in which the metal is shaped between two dies in such a manner as to fill the depression of both sides in relief by displacement of the material. Cold Working: The deformation of metal significantly below its recrystallization temperature such that work hardening occurs. Cupping: A press forming operation in which a cup-shaped part is produced from sheet metal. Cupping is often the first operation in the production of a complex deep-drawn part. Cup Test: A formability test in which a sheet blank is drawn or stretched into a hemispherical or cylindrical shape. Variations in blank diameter, cup height or forming load are utilized to evaluate sheet formability. (See also: Swift Cup Test, Fukui Cup Test) Cut-Off Die: Sometimes called a trimming die. The cut-off die can be the last die in a set of transfer dies which cuts the part loose from the scrap, or it can be a die which cuts straight sided blanks from a coil for later use in a draw die. Deep Drawing: A drawing operation where a part is produced from a blank by the action of a punch in which the sheet is

pulled into a die cavity and the flange of the blank is compressed in the circumferential direction. The area directly under the punch remains undeformed. Developed Blank: A flat blank with a shape that will produce a finished part with the desired configuration with a minimum of trimming operations. Die: (a) A complete tool used in a press for any operation or series of operations such as forming, impressing, piercing, and cutting. The upper member or members are attached to the slide (or slides) of the press and the lower member is clamped or bolted to the bed or bolster with the die members being so shaped as to cut or form the material placed between them when the press makes a stroke. (b) The female part of a complete die assembly as described in (a). Die Clearance: The space, on each side, between punch and die. Die Coating: Hard metal incorporated into the working surface of a die to protect the working surface or to separate the sheet metal surface from direct contact with the basic die material. Hard chromium plating is an example. Die Entry Radius: The radius at the edge of the die over which the work is drawn. Drawability: The ability of a sheet to form cup-shaped parts by the action of a punch under conditions that cause the flange or a portion of the flange to be drawn into the die cavity. Draw Bead: A ridge constructed around a portion of a die cavity to partially restrain metal flow. A groove in the mating blankholder allows die closing. Sometimes called Die Bead. Drawing: A sheet metal deformation process in which plastic flow results in a positive strain (e1) in one direction in the plane of the sheet surface and a negative strain (e2) at 90 to (e1) in the sheet surface. Drawing can only occur when sheet metal flow under the blankholder is permitted. The term drawing is sometimes loosely used to describe a wide variety of press forming operations. Drawing Ratio: The ratio of the blank diameter to the punch diameter. Draw Ring: A ring-shaped part of the die, integral or separate, over the inner edge of which the metal is drawn by the punch. (Also called Die Ring) Elastic Limit: The maximum stress to which a material may be subjected without any permanent strain remaining upon complete release of stress. Embossing: Displacing of metal a minor amount without noticeable reduction in sheet metal thickness.

Engineering Strain: Preferably called Nominal Strain. The unit elongation given by the change in length divided by the original length. Engineering Stress: Preferably called Nominal Stress. The unit force obtained when the applied load is divided by the original cross-sectional area. Erichsen Test: A test in which a piece of sheet metal, restrained except at the center, is deformed by a spherical ended plunger until fracture occurs. The height of the cup in millimeters at fracture is a measure of the ductility. Etching: Production of designs, including grids, on metal surface by a corrosive reagent or electrolytic action. Extruded Hole: A hole formed by a punch which first cleanly cuts a hole and then is pushed farther through to form a flange with an enlargement of the original hole. This may be a two-step operation. First Draw: The first drawing operation in a series which is usually performed on a flat blank. Flanging: A bending operation in which a narrow strip at the edge of a sheet is bent down (up) along a straight or curved line. It is used for edge strengthening, appearance, rigidity and the removal of sheared edges. A flange is often used as a fastening surface. Flow Curve: A graphical representation of the relationship between load and deformation during plastic deformation. Formability: The ability of sheet metal to be changed into a useful shape. (See Bendability, Stretchability, Drawability.) Formability Parameters: Material parameters that can be used to predict the ability of sheet metal to be formed into a useful shape. Forming Limit Diagram: An empirical curve showing the biaxial strain levels beyond which failure may occur in sheet metal forming. The strains are given in terms of major and minor strains measured from deformed circles, previously imprinted into the undeformed sheet. Fracture Load: The load at which splitting occurs. Fracture Strain ( f ): The true strain at fracture defined by the relationship:

=In

(initial cross-section area) (final cross-section area)

Fracture Strength: The engineering stress at fracture defined as the load at fracture divided by the original cross-section area. The fracture strength is synonymous with the breaking strength. Fracture Stress: The true stress at fracture which is the load for fracture divided by the final cross-section area. Free Bending: A bending operation in which the sheet metal is clamped at one end and wrapped around a radius pin. No tensile force is exerted on the ends of the sheet. Fukui Cup Test: A cupping test combining stretchability and drawability in which a round-nosed punch draws a circular blank into a conical shaped die until fracture occurs at the nose. Various parameters from the test are used as the criterion of formability. Gridding: Imprinting an array of repetitive geometrical patterns on a sheet prior to forming for subsequent determination of deformation. Imprinting techniques include: 1. Electrochemical marking (also called electro-chemcial or electrolytic etching) a grid imprinting technique utilizing electrical current, an electrolyte, and an electrical stencil to etch the grid pattern into the blank surface. A contrasting oxide usually is redeposited simultaneously into the grid. 2. Photoprinting a technique in which a proprietary photosensitive emulsion is applied to the blank surface, a negative of the grid pattern is placed in contact with the blank and the pattern is transferred to the sheet by a standard photographic printing practice. 3. Ink stamping. 4. Lithographing. Groove: Mating depression for the drawbead. Guided Bend Test: A test in which the specimen is bent to a definite inside radius of a jig. Hardness Test: A test to measure the resistance to indentation of a material. Test for sheet metal include Rockwell, Rockwell Superficial, Tukon and Vickers. Hole Expansion: Forcing the circumference of a punch hole to expand in a stretching mode. Hole Expansion Test: A formability test in which a tapered punch is forced through a punched or a drilled and reamed hole forcing the metal in the periphery of the hole to expand in a stretching mode until fracture occurs. Increase in Area: An indicator of sheet metal forming severity based upon percentage increase in surface area measured after forming. Ironing: Forming between dies whose side-wall clearance is less than the sheet metal thickness such that a thinning action occurs. This usually takes place in the side walls of deep-drawn parts where the inflow of

metal from the flange tends to thicken or wrinkle; subsequent ironing produces a more uniform wall thickness. Isotropy: A term indicating equal physical or mechanical properties in all directions. Keeler-Goodwin Diagram: The formability limit diagram for low-carbon steel commonly used for sheet metal forming. Limiting Drawing Ratio: An expression of drawability given by the highest drawing ratio attained in a series of tests such as the Swift Cupping Test. Lock Bead: A ridge constructed around a die cavity to completely restrict metal flow into the die. Lubricant: Any substance interposed between two surfaces in relative motion for the purpose of reducing the friction and/or wear between them. Lders Lines or Lders Bands: The surface strain markings associated with discontinuous yielding that occur as a result of deformation operations such as press forming. Lders lines or bands are also referred to as stretcher strains. Lders Strain: Surface markings, usually of a wedge-shaped pattern, which appear in early stages of plastic elongation of soft, annealed-last steel or other steels where the interstitial elements carbon and nitrogen have not been stabilized. Major Strain: Largest principal strain in the sheet surface. Often measured from the major axis of the ellipse resulting from deformation of a circular grid. Michrohardenss Test: An indentation test using diamond indentors at very low loads usually in the range of 1 to 1000 grams. Minimum Bend Radius: The smallest radius about which a metal can be bent without exhibiting fracture. It is often described in terms of multiples of sheet thickness. Minor Strain: The principal strain in the sheet surface in a direction perpendicular to the major strain. Often measured from the minor axis of the ellipse resulting from deformation of a circular grid. Necking: Localized thinning that occurs during sheet metal forming prior to fracture. The onset of localized necking is dependent upon the stress state which is affected by geometrical factors. Nominal Strain: The unit elongation given by the change in length divided by the original length. Also called Engineering Strain.

Nominal Stress: The unit force obtained when the applied load is divided by the original cross-sectional area. Also called Engineering Stress. "n" Value: A term commonly referred to as work-hardening exponent derived from the relationship between true stress and true strain given by = Kn. Often called work strengthening exponent and in Mechanical Engineering nomenclature may be called m. Normal Anisotropy: A condition in which a property or properties in the sheet thickness direction differ in magnitude from the same property or properties in the plane of the sheet. Offal: Sheet metal section trimmed or removed from the sheet during the production of shaped blanks or the formed part. Offal is frequently used as stock for the production of small parts. Olsen Test: A formability test in which a piece of sheet metal, restrained except at the center, is deformed by a standard steel ball until fracture occurs. The height of the cup in thousandths of an inch at time of failure is a measure of the ductility. Piercing: Forming a hole in a sheet metal with a pointed punch with no metal slug fallout. Planar Anisotropy: A term indicating variation in one or more physical or mechanical properties with direction in the plane of the sheet. The planar variation in plastic strain ratio is commonly designated as >r, given by: >r

= (ro + r90 2r45)/2

The earing tendency of a sheet is related to r. As r increases so the tendency to form ears increases. Plastic Instability: The deformation stage during which plastic flow is nonuniform and necking occurs. Plastic Strain Ratio, (r): A measure of normal plastic anisotropy is defined by the ratio of the true width strain to the true thickness strain in a tensile test. The average plastic strain ratio rm is determined from tensile samples taken in at least three directions from the sheet rolling direction, usually at 0, 45, and 90. rm is calculated as follows: rm = (ro + 2r45 + r90)/4 Principal Strain: The normal strain on any of three mutually perpendicular planes on which no shear strains are present. Progressive Die: A die planned to accomplish a sequence of operations as the strip or sheet of material is advanced from station to station, manually or mechanically. Progressive Forming: Sequential metalworking at consecutive stations either with a single die or with separate dies.

Proportional Limit: The stress above which stress is no longer proportional to strain, i.e., Hookes Law ceases to apply (see Elastic Limit). Punch: The member of a tool that forces the metal into the die during blanking, coining, drawing, embossing, forging, stretching, or similar operations. Punch Nose radius: The shape of the punch end, contacting the material being formed to allow proper material flow or movement. Quench Aging: Hardening by precipitation which results after the rapid cooling from solid solution to a temperature below which the elements of a second phase become supersaturated. Precipitation occurs after the application of higher temperatures and/or times and causes increases in yield strength, tensile strength, and hardness. r Value: The ratio of true width strain to true thickness strain. Often called Plastic Strain Ratio. Reverse Redrawing: An operation after the first drawing operation in which the part is turned inside out by inverting and redrawing, usually in another die, to a smaller diameter. Shear Burr: A raised edge resulting from metal flow induced by blanking, cutting, or punching. Shearing: A cutting operation in which the work metal is placed between a stationary lower blade and movable upper blade and severed by bringing the blades together. Cutting occurs by a combination of metal penetration and actual fracture of the metal. Slitting: The cutting of lengths of sheet material into narrower lengths by means of one or more pairs of circular knives. Springback: Elastic recovery of a portion of the deformation produced during forming. It results in a difference in shape between the part and the die. Stamping: A general term to denote all pressworking. In a more specific sense stamping is used to imprint letters, numerals, and trademarks in sheet metal, machined parts, forgings, and castings. A tool called a Stamp with the letter or number raised on its surface, is hammered or forced into the metal leaving a depression on the surface in the form of the letter or number. Stiffness: Resistance to elastic deformation. Strain Aging: A phenomenon that occurs in some materials following plastic deformation. In low-carbon steel sheet,

strain aging results in a return of discontinuous yielding, an increase in yield strength and hardness, and a decrease in ductility without a substantial change in tensile strength. Strain Lines: Surface defects in the form of shallow line type depressions appearing in sheet metals after stretching the surface a few percent of unit area or length. (See also Lders Lines.) Strain Rate Sensitivity: The degree to which mechanical properties are affected by changes in deformation rate. Strength: The ability of a material to withstand an applied force. Strength Coefficient (K): A constant related to the tensile strength used in the power law equation = Kn. In Mechanical engineering nomenclature it is called o and the power law equation is given = on. See also n value. Stretchability: The ability of a material to undergo stretch-type deformation. Stretcher Strains: Strain lines developed by a stretching operation in a sheet metal. See also Lder Strain, Lders Lines, Strain Lines and Yield Point Elongation. Stretch Forming: A process in which a sheet section is formed over a block of the required shape while the blank is held in tension. Stretching: A mode of deformation in which a positive strain is generated on the sheet surface by the application of a tensile stress. Surface Hardness: The hardness of that portion of the material very near the surface as measured by micro-hardness or superficial hardness testers. Surface Roughness: The fine irregularities in the surface texture which result from the production process. Considered as vertical deviations from the nominal or average plane of the surface. Surface Texture: Repetitive or random deviations from the nominal surface which form the pattern of the surface. Includes roughness, waviness, and flaws. Swift Cup Test: A test for deep drawability using circular blanks of various diameters which are drawn into a flat or hemispherical bottomed cup. Drawability is assessed by the Limiting Drawing Ratio. Tearing: Failure and localized separation of a sheet metal. Tensile Ratio: The ratio of the tensile strength to yield strength. It is the inverse of the yield ratio.

Tensile Strength or Ultimate Tensile Strength: The strength calculated at the maximum load, in a tensile test, by dividing the maximum load by the original cross-sectional area. Total Elongation (et): A parameter measured in a tensile test used as a measure of ductility defined by:
Final gage length original gage length x 100 original gage length

Work Hardening: An increase in hardness and strength caused by plastic deformation at temperatures lower than the recrystallization temperature. Sometimes referred to as strain hardening. Work Hardening Exponent (n): The exponent in the relationship = K n where is the true stress and is the true strain and K is the constant commonly called the Strength Coefficient. In some Mechanical Engineering nomenclature the strain hardening exponent is defined from Myers equation where L = Adn L is the applied load in an indentation test and d is the diameter of the indentation. Yield Point Elongation: The extension associated with discontinuous yielding which occurs at approximately constant load following the onset of plastic flow. It is associated with the propagation of Lders lines or bands. Yield Stress: A stress at which a steel exhibits the first measurable permanent plastic deformation. Yield Ratio: The ratio of the yield strength to the tensile strength. It is the inverse of the tensile ratio. Youngs Modulus or Elastic Modulus: A measure of the rigidity of a metal. It is the ratio of stress, within the proportional limit, to corresponding strain. Youngs Modulus specifically is the modulus obtained in tension or compression.

You might also like