You are on page 1of 103

LOADS AND CONSTRAINTS

Beam Preload
A beam preload creates an axial load over the length of a beam element. This is useful for
simulating fasteners, wire rope, and other loads that are internal to the model or part (as
opposed to an external load such as a nodal force). Unlike external loads which are constant
throughout the analysis, the preload magnitude is the magnitude in the element if the rest of
the structure were infinitely stiff. Since the structure is not infinitely stiff, one result of a
preload is that the structure deforms and relieves a portion of the preload. See Practical
Considerations below.

Apply Preload Loads


In linear analyses, beam preloads are available only for static stress. They are not available
for Natural Frequency (Modal) with Load Stiffening nor Critical Buckling Load.
In nonlinear analyses, beam preloads are available for MES and for Static Stress with
Nonlinear Material Models. It is also possible to apply a beam preload in a Natural Frequency
(Modal) with Nonlinear Material Models analysis. However, it will produce no effect on the
results, since this analysis type does not account for load stiffening effects. Beam preloads are
not available for an MES Riks analysis.
1. Define the Element Type for the part to be Beam elements.
2. Select one or more beam elements using the Selection
command.

Select

Lines

3. Right-click in the display area and choose the Add pull-out menu and
select the Beam Preload command to add a preload load each beam
element.
4. Enter the magnitude of the preload in the Axial Force field. A positive
value indicates the element is initially in tension, so the analysis draws the
beam ends together. A negative value indicates the element is initially in
compression, so the analysis moves the beam ends apart.
5. A beam element with a preload has the symbol B on the line segment. To
modify the preload, either select the preload entry in the FEA Object
Groups branch of the tree view, or select the B symbol in the display area
while in the line selection mode (Selection Select Lines).
Note:

In addition to the context menu command, you can also click the Beam
Preload command within the Beam Loads panel of the ribbon Setup
tab.

The initial strain is considered to be zero. Thus, the calculated strain does
not appear to agree with the calculated stress unless you consider the
prestrain.

For linear analyses, the preload is applied to all load cases in the analysis.
It is not affected by any load multipliers.

For plastic material models in a nonlinear analysis, the pre-strain is


calculated, so the effect appears in the strain results.

Practical Considerations
As noted above, some of the preload applied to the beam element is relieved in the analysis
because the members compress in response to the load. If the preload provided is intended to
be the final load, such as the preload in a bolt due torquing it with a torque wrench, then the
beam preload applied must be increased to compensate for the compression of the parts. If the
stiffness of the members and the final load were known, the initial preload to apply to the
beam elements could be calculated as follows:

Consider a beam (represented by red spring K b ) that is stretch to create a preload of amount
P, and then attached to the members (represented by blue spring K m ). The beam and member
then compress an amount . Since the beam is stretched before attaching to and compressing
the member, the final load F' in the beam equals the preload minus the stiffness times the
compression, while the equal (but opposite) load in the member equals the stiffness times the
compression. Thus, there are two equations and two unknowns (F' and ).
Final load in beam F' = P - Kb
Final load in member F' = Km
Solving these two equations for the final load F' gives
F' = P[Km/(K m +Kb)]

where Kb is the stiffness of the beam, Km is the stiffness of the member, and P is the preload
applied to the beam.
For a bolted connection, there are many formulas for calculating the stiffness of the members
depending on the assumptions about the pressure distribution under the bolt. One such
formula is as follows which considers an apex angle of 30 degrees for the pressure
distribution and a washer diameter of 1.5 times the bolt diameter:

where L is the total thickness of the bolted members and d is the diameter of the bolt and
hole. (Reference: Budynas, Richard G and Nisbett, J. Keith, Shigley's Mechanical
Engineering Design, McGraw-Hill, Inc, Eighth Edition)
The item of interest for this topic is the ratio P/F'. In many cases, the known preload is the
final load in the bolt and member (F'), so the above formulas can be used to calculate the
preload P to apply to the beam elements. The following table gives several results. For
example, when L/d is 2, the preload to apply to the beam elements is 1.35 times the final
preload in the bolt and members. (Obviously, each design is different, so the above equation
for Km and the table are only representative of the approach.)

0.50
1.00
2.00
3.00
4.00
6.00
10.0

Notes:

If the members are composed of several materials, the stiffness can be


calculated as springs in series. Recall that the equivalent stiffness K eqv in
this situation is found from

1/Keqv = 1/K1 + 1/K2 + 1/K3.

If one material is much weaker than the others, such as a soft gasket
material, then the softer material governs and K eqv Kgasket.

In some critical cases, it may be necessary to perform the analysis with no


loads applied to the model except for the preload to determine what
preload P is required to get the final load F'.

Similar approaches can be used when using temperature to apply a preload to part of the
model.
Bearing Loads

A bearing load can only be applied to a cylindrical surface. The surface does not need to be a
full 360 cylinder but can be any portion of a cylinder. In addition, the surface can belong to a
CAD-based model or a hand-built, structured mesh. A bearing load represents the typical
compressive load distribution that occurs in contact areas between shafts and bearings or
bushings.

Characteristics of Bearing Loads:

Bearing loads can consist of a radial load, a thrust load (axial direction), or
both.

Bearing loads are converted to nodal forces. The effect of this conversion
can be seen in the Results environment. Forces at edge and corner nodes
are noticeably smaller than forces on nearby nodes that are within the
interior of the surface. This behavior is normal because edge and corner
loads are associated with a smaller element facial area than interior nodes
are.

Thrust loads are distributed along all nodes on the selected cylindrical
surface, as shown below.

Radial loads are applied in a parabolic distribution and are always confined
to a maximum cylindrical load zone of +/-90 from the radial vector
direction. Consider the images below, which explain how the radial load is
distributed in different situations:

(a) Radial load on a full 360 cylindrical surface.

Important: Regardless of the number of partial surfaces included in a


single bearing load, the Magnitude specified in the dialog box is the
total load magnitude. This behavior differs from other Autodesk
Simulation loads, where the specified magnitude is applied to each
selected entity. For example, if a cylinder is divided into four 90 surfaces
and a 1,000 pound bearing load is defined with all four surfaces selected,
the total load will be 1,000 pounds. Conversely, when a 1,000 pound
surface force is applied, the total load is 4,000 pounds (1,000 lbs./surface
x 4 surfaces).

The axis orientation of the cylindrical surface is calculated automatically.


The thrust vector direction cannot be manually specified, since it is always
in the axial direction. However, it can be reversed by clicking the Toggle
Vector Direction button.

The radial load direction can be specified in three different ways:


1. Use X, Y, or Z radio buttons to specify a global direction.
2. Manually enter custom vector components in the X, Y, and Z fields.
3. Click the Radial Vector Selector button and then click two points
on the model to indicate the direction graphically.

If the specified vector is skewed relative to the purely radial direction of the surface, it
is automatically adjusted to align with the purely radial direction.

A bearing load may act against an internal surface (such as the inside of a
bearing bore), or against an outside surface (such as the shaft extension
where a bearing is to be located). The determination of internal versus
external loads occurs automatically for solid elements, as shown in the
following image:

Notice that the vector direction is the same for both parts. The selected surface
determines whether the load is internal or external. Both loads are applied in the +Y
direction for this example. The left load is applied to the inner surface of the bushing.
The right load is applied to the outer surface.

For planar elements (plates/shells), you must specify whether the load is
to be applied to the inside or outside face. The program cannot infer the
intended load from the vector alone. When you apply a bearing load to a
planar element, the dialog box will contain an additional input field (Load
applied to), which is not visible for other element types. Consider the
following plate/shell example:

Again, the vector direction is the same for both parts (+Y). The left load occurs when
the Inside Face option is selected. The right load occurs when the Outside Face
option is selected.

Note: The specification of inside or outside face, in the Bearing Load


Object dialog box, is independent of the Element Normal Point, which is
used to orient planar elements for other types of loads.

To Apply a Bearing Load:


1. If you have surfaces selected, you can right-click in the display area and
select the Add pull-out menu. Choose the Surface Bearing Load
command. You can also access this command via the ribbon (
Setup
Loads
Bearing). You can invoke this command from the ribbon either
prior to or after selecting the application surface.
2. In the Load Case / Load Curve field, specify the load case number (static
analysis) or load curve number (transient stress or nonlinear analysis) that
will control the bearing load over time.
3. Specify the thrust force and/or radial force in the Magnitude fields.
4. Click the Toggle Vector Direction button if you wish to reverse the
thrust load direction.
5. Choose one of the global X, Y, or Z radio buttons to specify the radial load
direction; or...
o

Activate the Custom radio button and manually enter the X, Y, and
Z vector components in the X, Y, and Z input fields; or...

Click the Radial Vector Selector button and then click two points
on the model to indicate the radial load direction.

Whether manually or graphically specified, the vector is automatically adjusted, if


necessary, to conform to the purely radial direction of the selected cylindrical surface.
6. For planar element (such as plates or shells), specify whether to apply the
load to the Inside Face or Outside Face using the Load applied to
menu button.
7. Optionally, type a Description of the load in the provided field.
8. Click OK to apply the load or Cancel to abort the command.
Centrifugal Load
Note: The information in this section applies to all linear and nonlinear
analyses that support centrifugal loads.

A centrifugal load simulates the effect of the entire model spinning about an axis you specify.
Only parts with a nonzero material mass density are affected. The model does not actually
experience rotation. Rather, the equivalent forces that would occur as a result of angular
rotation and/or angular acceleration are calculated and applied to the nodes of each element.

Linear Analyses
Three types of linear analysis support centrifugal loadsStatic Stress with Linear Material
Models, Natural Frequency (Modal) with Load Stiffening, and Critical Buckling. The model
can be spinning at a constant rate and/or can be undergoing a constant angular acceleration
rate.
The command implementation is slightly different among the linear analyses. In all three
cases, the actual load is defined within the Centrifugal tab of the Analysis Parameters dialog
box, and the axes of rotation and angular acceleration can pass through any point is 3D space
(that is, the axes do not have to pass through the global coordinate origin). The differences
are as follows:

Static Stress with Linear Material Models:


o

The Centrifugal load can be accessed in three ways...


1. Right-click the Centrifugal heading under the Analysis Type
heading in the browser (tree view) and choose the Edit
command.
2. Use the ribbon command, Setup

Loads

Centrifugal.

3. Access the Analysis Parameters dialog box (either via the


browser or ribbon) and click on the Centrifugal tab.

Global load case multipliers separately control whether or not the


Angular Velocity (Omega) and/or Angular Acceleration (Alpha)
loads, as defined within the Centrifugal tab, are active for each load
case.

The axis of angular velocity and the axis of angular acceleration can
be separately defined (and can be the same or different).

The axes or angular velocity and angular acceleration can be along


any line in 3D space, as specified via direction vectors.

Natural Frequency (Modal) with Load Stiffening and Critical Buckling:


o

The Centrifugal load can be accessed in two ways...


1. Use the ribbon command, Setup

Loads

Centrifugal.

2. Access the Analysis Parameters dialog box (either via the


browser or ribbon) and click on the Centrifugal tab.
o

There are no global load case multipliers for the centrifugal load.
Rather, it is enabled solely via the Include specified centrifugal
load checkbox within the Centrifugal tab of the Analysis
Parameters dialog box. If enabled, the centrifugal load affects every
resultant mode shape, modal frequency, and buckling multiplier.

The axis of rotation and angular acceleration must be the same (not
separately specified).

The axis of rotation and angular acceleration must be one of the


three global axes (X, Y, or Z).

Nonlinear Analyses
Two types of nonlinear analysis support centrifugal loadsMES and Static Stress with
Nonlinear Material Models. Load curves control both the rotation speed and the angular
acceleration rate over time. The same load curve can be used for both, or you can specify two
different load curves for rotation and angular acceleration. The centrifugal load is defined
within the Centrifugal tab of the Advanced Analysis Parameters dialog box and can be
accessed using one of the following two methods:
1. Use the ribbon command, Setup

Loads

Centrifugal.

2. Access the Analysis Parameters dialog box (either via the browser or
ribbon), click the Advanced button, and click on the Centrifugal tab.

For either type of nonlinear analysis that supports centrifugal loads...

The axis of rotation and the axis of angular acceleration can be separately
defined (and can be the same or different).

The axes of rotation and angular acceleration can pass through any point
is 3D space (that is, the axes do not have to pass through the global
coordinate origin).

The axes or rotation and angular acceleration can be along any line in 3D
space, as specified via direction vectors.

Apply Centrifugal Loads


To apply a centrifugal load to a model...
1. Access the appropriate dialog box using one of the methods described
above (differs among the various analysis types).
2. If included within the dialog box, activate the Include specified
centrifugal load checkbox.
3. Specify the appropriate magnitudes under Angular Velocity (Omega)
and Angular Acceleration (Alpha).
4. For nonlinear analyses, select the load curve numbers to use for angular
velocity and angular acceleration.
5. Specify the axis orientation using one of the methods described above
(differs among the various analysis types). For some analysis types, the
rotation and angular acceleration axes may be different, and for others,
the same axis must be used for both (as detailed above). In addition, some
analysis types support non-global axis orientations and others do not (also

as detailed above). Use the Point on Axis coordinates and the Direction
coordinates or pull-down global axis list to specify the location and
direction of the axis (or axes) . The angular rotation follows the right-hand
rule about the vector you define.
6. For linear static stress analyses, set the Omega and/or Alpha global load
case multipliers within the Multipliers tab of the Analysis Parameters
dialog box.
Tip: A nonzero multiplier is required to activate the load. See the page
Analysis Parameters: Static Stress with Linear Material Models.

For example, take the circle centered at the origin in Figure 1. To specify rotation about the Z
axis at the origin, enter Point on Axis coordinates of (0,0,0) and Direction coordinates of
(0,0,1) - the circle rotates about its center and the center of the circle remains stationary.

Figure 1: Model 1 Rotating About Origin


However, if you have the circle as shown in Figure 2 centered at (3,0,0) with the same
settings for the centrifugal load, the load is based on rotation of the circle about the origin. In
other words, the center of the circle moves in a circular path about the global origin. Figure 3
shows a graphical depiction of the path of the circle from Figure 2 under the specified
centrifugal load.

Figure 2: Model 2 Circle with its Center Not at the Global Origin

Figure 3: Interpretation of Model 2 Rotating About Origin


Note: For all three linear analyses that support centrifugal loads, remember that,
despite the simulated motion effects, these are still static analyses. Therefore,
the model needs to be statically stable in all directions. Typically, a rotating part
or assembly is attached to a shaft or lever. Two common constraint schemes are
as described below:

Fully-fix the area or areas of attachment to the supporting shaft or lever.

Alternatively, constrain tangential motion along the surface of the hole


where the rotating parts are fit to the shaft or lever. This constraint takes
care of two of the three translational degrees of freedom. In addition,
constrain a point, edge, or surface against translation in the axial
direction. For example, constrain the area of contact with the shaft
shoulder against translation in the normal direction, which prevents axial
motion under thrust loads and takes care of the third translational DOF.

Distributed Loads

A distributed load is a load that is applied over the length of a beam element. A distributed
load can be applied in any direction specified by a vector.

In the case of a Linear analysis, for the distributed load to be applied to your model, you
must assign a Pressure multiplier in the Multipliers tab of the Analysis Parameters dialog
box. The product of the Magnitude and the Pressure multiplier updates the magnitude of the
distributed load.
In the case of a Nonlinear analysis, a load curve controls the magnitude of the distributed
load during the event duration. You specify the load curve number within the Creating Beam
Distributed Load Object dialog box.

Apply Distributed Loads


If you select one or more beam elements using the Selection Select Lines command and
right-click in the display area, you can select the Add pull-out menu and select the Beam
Distributed Loads command to add a distributed load to each beam element.
If the magnitude and direction of the load is constant across the length of each selected beam,
specify the magnitude of the load per unit of length in the Magnitude field and the direction
of the load in the Direction section. Click the Flip Direction button (
) to invert the sign of
the applied load, reversing its direction. If the magnitude or direction of the load varies across
the length of each selected beam, deactivate the Uniform check box and specify the
magnitude of the force per unit of length at each end in the I-Node Magnitude and J-Node
Magnitude fields and the direction in the I-Node Direction and J-Node Direction sections.
The magnitude of the load varies linearly between the nodes. This is done on a per-beam
basis.
If you are applying a non-uniform load to the beam, you can have the load start and end at
any point along the beam. This is done by specifying a ratio of the length of the beam in two
ratio from I-Node fields. The first field determines how far from the I-Node the load starts.
The second field determines how far from the I-Node the load ends. For example, if you want
a load to be applied only along the third-quarter of a beam, you specify 0.5 in the first field
and 0.75 in the second field.
Tip: The axis 1 orientation of the beam elements can be displayed using the
View
Visibility
Object Visibility
Element Axis 1 command. This axis
lies along the direction of the beam element line and the positive direction is
from the I-Node to the J-Node. If axis 1 needs to be reversed for some elements,
this can be done by selecting the elements ( Selection
Select
Lines),
right-clicking, and choosing Beam Orientations
Invert I and J Nodes.

For Nonlinear Analyses Only


Specify the load curve that will be used to multiply the distributed load in the Load Case /
Load Curve field. Press the Curve button to define a load curve in the Load Curve Editor, or
use the Analysis: Parameters dialog box. Specify an additional multiplier applied to this
load in the Multiplier field.
For the load to maintain the same orientation with respect to the element as it deforms,
activate the Follows Displacement check box (and the Large Displacement analysis type in

the Element Definition). When using Follows Displacement, the following additional
calculations are performed:
1. The user's load is converted from global X, Y, Z directions into the beam
element's local 1, 2, 3 directions at the start of the analysis.
2. Throughout the analysis, the loads remain in the local 1, 2, 3 directions.
So, as the beam moves and rotates, the local axes also change orientation
in space. Consequently, the direction of the loads in global X, Y, Z are
changing.
3. The magnitude of the load remains constant. But if the length of the
element changes, the total load changes. If the element gets twice as
long, the load will be twice as much (constant magnitude x updated
length).
Note: The visualization of distributed load in the Results environment may not
follow the beams properly when the beams experience axial rotations. This is just
a visualization issue; the calculated results are not affected.
Forces

Force loads can be applied to nodes, edges, or surfaces.

Force loads can be applied in any direction, as specified by a vector.


Additionally, a surface force can be applied normal to the surface.
Depending upon the type of model (CAD-based or hand-built), Force loads
may or may not be available. Depending upon the specific type of force
load (nodal, edge, or surface), local coordinate systems may or may not
be supported. See the following bullets for details.
o

Nodal forces are supported for all CAD-based and hand-built


models. Local coordinate systems are supported for nodal force
objects.

Edge forces apply nodal forces to each node along the edge. The
magnitudes of the nodal forces are calculated so that the force is
evenly distributed along the edge. Edge forces are applicable only
to CAD-based solid or surface models and 2D Mesh Generation
models. Local coordinate systems are supported for edge forces.

Surface forces are only applicable to CAD-based solid or surface


models. A surface force applies an equivalent pressure to the
selected surface, so the distribution of force is uniform even if the
nodes are unevenly spaced. Local coordinate systems are NOT
supported for surface forces. The load direction must either be
normal or specified using the global axis directions (radio buttons or
vector components).

As an alternative to applying the specified force magnitude to each of several


selected surfaces, you can choose to distribute a specified force magnitude
over two or more surfaces. If this option is activated, the total force is

distributed based on the relative area of each surface selected. See the Apply
Forces section below for more details.
What does a force do?
The data entered for the force is applied to each object selected and in the
direction you specify. Therefore, if you select 10 nodes and specify a 10 lb force
magnitude in the X direction, you have just applied 100 lbs (10 objects * 10 lbs
per object) to your model. The same is true if you apply the force to 10 surfaces
or 10 edges. The exception to this rule is for surface forces applied to two or
more surfaces when the Distribute magnitude across all surfaces option is
enabled. See the Apply Forces section below for more details.

Analysis-Specific Parameters:
1. For Linear analyses:
o

Nodal and edge forces are applied to a load case. The load case is
defined in the Load Case/ Load Curve field. If you want a nodal or
edge force to be applied in multiple load cases, you can copy it to a
new load set and change the value in the Load Case/ Load Curve
field.

For a surface force to be applied to a model a Pressure multiplier


must be defined in the Multipliers tab of the Analysis
Parameters dialog box.

2. For Nonlinear analyses:


o

All forces will follow a load curve throughout the analysis. Select the
loads curve in the Load Case/ Load Curve field.

Apply Forces
If you have nodes, edges or surface selected, you can right-click in the display area and select
the Add pull-out menu. Select the Nodal Force, Edge Force or Surface Force command.
You can also access this command from the ribbon (Setup Loads Force). You can click
the ribbon command either before or after selecting the objects to which you wish to apply
the force.

Specify the magnitude of the force that is applied to each selected object
in the Magnitude field. Alternatively, specify the total force to be
distributed over multiple surfaces in the Magnitude field.

Specify the direction of the load in the Direction section.

Click the Flip Direction button (


reversing its direction.

If two or more surfaces are selected, the Distribute magnitude across


all surfaces option will be available. When this option is activated, the

) to invert the sign of the applied load,

specified magnitude is distributed on a pro rata basis depending upon the


area of each selected surface. The example in the following table explains
how the load is distributed:

Example:
Assume that a total force of 2,500 N is distributed over three surfaces (a, b, and c)
with areas of 100, 300, and 600 mm2, respectively.
Surface
a
b
c

Caution: Be careful when running a Parametric Study on models with a


distributed surface force applied. The forces applied to individual surfaces
of a distributed force group are not updated when the model geometry
changes. If any of the parameters you are manipulating affect the area of
a surface to which a distributed force is applied, the force distribution will
no longer be correct.

The arrows used to indicate applied forces point in the specified load vector direction,
including the Normal direction for forces applied normal to surfaces. However, the lengths of
the force arrows are not scaled to indicate the relative magnitude of the applied forces.
For linear analyses, if you are applying an edge or nodal force, specify the load case or load
curve in which you want the force placed in the Load Case/Load Curve field. Surface forces
are controlled by the global load case Pressure multiplier within the Analysis Parameters
dialog box. For nonlinear analyses, specify the load curve number within the Load
Case/Load Curve field for all surface, edge, and nodal forces.
Note: For information about how nodal loads are applied at duplicate vertices,
see the comments under the Application of Loads and Constraints at Duplicate
Vertices heading on the Loads and Constraints page.

Import Reaction Forces From an Electrostatic Analysis


The structural reactions to the electrostatic force between conductors (usually significant only
in MEMS applications) can be calculated. If an electrostatic analysis has been performed in a
different model, and if the reaction forces were calculated in the different model, then those
reaction forces can be applied to the structural model. To activate the output of the reaction
forces, see the Calculating the Forces and Charge Caused by the Electrostatic Field section
on the Analysis Parameters: Electrostatic Field Strength and Voltage page. (Although the

electrostatic forces in the binary results file are in units of voltage current time/length, the
software automatically converts this to force units when applying the loads to the model.)
Attention: Due to changes in the software, electrostatic forces calculated prior to
V23.1 (February 2009) should not be used with software versions V23.1 or newer.
If these results are needed in V23.1 or newer, the analysis must be performed
again.

The nodes in the stress model need to match the nodes in the electrostatic model where the
reaction forces were calculated. But, the electrostatic parts do not need to be included in the
stress analysis. For example, the force may be calculated on the surface of a part representing
the dielectric material between conductors (such as air). Air between conductors would not be
included in the stress analysis.
The method to apply the reaction forces to the stress model is as follows:
1. With nothing selected, right-click in the display area of the FEA Editor.
Select the Loads from File command.
2. Press the Browse button in the Results File column. A dialog box
appears. Use the Files of type: pull-down to select the Electrostatic
Reaction Forces (*.efr) type of file to read for the loads.
3. Select the file and click Open.
4. Since electrostatic analyses can only have a single load case, the Load
Case from File column will show zero (0). However, you can import
multiple sets of loads into a single model by pressing the Add Row
button. The loads assigned to the other rows of the spreadsheet can be
from the same file and design scenario, from different files, or from
different design scenarios. For example, you may want to apply the same
electrostatic forces to more than one load case in the stress analysis. Or,
you may want to calculate stresses for electrostatic force results at two
different applied voltages, each one being from a different design
scenario.
5. The loads are placed in a specific load case or load curve in the stress
analysis; enter the load case/load curve in the Structural Load Case
field.
6. If you want the loads to be multiplied by a constant value before being
applied to the model, specify the constant value in the Multiplier column.
7. Press the OK button.
Note:

The imported loads do not appear in the FEA Editor. They appear in the
Results environment after doing a Check Model or performing the analysis.

For linear analyses, don't forget to define the number of load cases as
necessary for the imported loads. If the model is set up for 2 load cases

and a load imported from an electrostatic analysis is assigned to load case


5, the load on load case 5 does not exist. The model has only two load
cases. To create the additional load cases in this example, add three rows
on the Multipliers tab of the Analysis Parameters dialog box.

For nonlinear analyses, don't forget to define the load curves as necessary
for the imported loads. If a load imported from an electrostatic analysis is
assigned to load curve 5 and this load curve is not defined, the multiplier
will be 0!

Tip:

The Loads from File dialog box can also be used for the following:

Load temperatures from another model. See the Temperatures page.

Loads from a text file can be imported to the structural analysis. See the
Loads from File page.

Gravity or Acceleration

A gravity or acceleration load applies an acceleration value to any part that has a mass density
defined. The acceleration can be applied along any direction.
In the case of Linear analyses (Static Stress, Modal with Load Stiffening, and Critical
Buckling), the acceleration value is constant. For Nonlinear analyses (MES and Static Stress
with Nonlinear Materials), the acceleration is controlled by a load curve and can be increased
gradually and/or otherwise varied over time.

Apply Gravity/Acceleration
To apply a gravity or acceleration load to a model, right-click the Gravity/Acceleration
heading under the Analysis Type heading in the tree view and select the Edit command. Note
that the text of this command will be gray before gravity is activated and defined. However, it
is not grayed-out. That is, the heading is still right-clickable and the Edit command is
available. The heading text becomes black once gravity is set up.
Note: You can also click the Gravity command within the Loads panel of the
ribbon Setup tab. Either method displays the Gravity/Acceleration tab of the
Analysis Parameters dialog box.

To apply the acceleration due to gravity on Earth, press the Set for standard gravity button.
The standard value for the acceleration of gravity is applied in the units of the model. To
apply a different acceleration magnitude, specify this in the Acceleration due to body force
field. Next, use the X multiplier, Y multiplier, and Z multiplier fields to define the vector
along which the acceleration is applied. Specifying a value in only one of these fields applies
the acceleration in that direction. Specifying values in more than one of these fields applies
the acceleration along an arbitrary vector. The value in the Acceleration due to body force
field is multiplied by the values in these three fields before it is applied to the model in that
direction.

The X, Y, and Z multiplier values are not normalized to a unit vector, as they are for many
other types of loads. Consider the following two examples:

If you wish to apply standard gravity at a 45 angle between the +X and


-Z directions, click the Set for standard gravity button, set the X
multiplier to 0.707107 and set the Z multiplier to -0.707107.
Tip: cos (45) = 0.707107

The resultant vector is 1.0 and is in the desired direction.

If you wish to simultaneously apply a 3g acceleration in the -Y direction


and a 5g acceleration in the -Z direction, click the Set for standard
gravity button, set the Y multiplier to -3, and set the Z multiplier to -5.

In the case of Nonlinear analyses, choose the load curve that will control the
gravity/acceleration load using the available pull-down list. The load curve must be
previously defined and the model saved in order for it to appear in this list.
When gravity is applied to the model, an arrow showing the direction of the gravity load
appears in the display area in both the FEA Editor and the Results environment. The gravity
arrow can be shown in the display area at either of these locations:

At the center of the box that encloses the model. If parts are hidden,
shown, deactivated, or activated; the location of the arrow updates when
the model is enclosed (View
Navigate
Enclose).

Super-imposed on the mini-axis. Use the


Options
Graphics
Mini-axis
Include gravity (when applied) check box to control where
the gravity arrow appears at.

In the FEA Editor, the gravity arrow itself can also be shown or hidden by right-clicking on
the Gravity/Acceleration entry in the tree view which is located underneath the Analysis
Type entry.
Hydrostatic Pressure

Hydrostatic pressure varies linearly from the level of the fluid in the direction of increasing
depth of the fluid. The magnitude of the hydrostatic pressure = (fluid density) x (depth below
the fluid surface).
You can apply hydrostatic pressure to plate, shell, brick, 3D kinematic, nonlinear membrane,
3D gasket, or tetrahedral elements. By default, the pressure is normal to the face of the
elements.

Apply Hydrostatic Pressures


If you have surfaces selected, you can right-click in the display area, select the Add pull-out
menu, and choose the Surface Hydrostatic Pressure command. This command is also
available on the ribbon (Setup Loads Hydrostatic Pressure).

To apply a hydrostatic pressure to any supported element type...

You must specify the weight density (in the Fluid Density field) of the
fluid that causes the hydrostatic pressure.

The hydrostatic pressure can increase along any direction. Specify a point
on the top of the fluid in the X, Y, and Z fields in the Point on Fluid
Surface section. Only elements below this point receive a pressure.

Using the Surface Normal of Fluid section, specify a vector that is


normal to the surface of the fluid and points into the fluid (that is, in the
direction of increasing fluid depth and gravity). When the surface normal is
aligned to a global axis, only one of the values for the Point on Fluid
Surface is critical.

For plate elements (linear analyses) and general shell elements (nonlinear
analyses), a Side selector is provided within the Creating Surface
Hydrostatic Pressure Object dialog box. Choose the Top or Bottom side
from the pull-down list. The bottom is the side facing the element normal
point, as defined within the Element Definition dialog box. For general
shell elements the additional two choices, Both and Neither, are also
available from the Side selector. Regardless of whether the load is applied
to the top or the bottom surface, the direction is towards the element.

If you are performing a transient stress (direct integration) analysis or a


nonlinear analysis, select the load curve that the pressure follows in the
Load Curve field.

Note: Software versions prior to 2013 included an additional Multiplier field for
hydrostatic loads in nonlinear analyses only. This option has been eliminated.
Essentially, all new loads will be based on a non-changeable multiplier of 1. For
older models opened in version 2013 or later, the specified fluid density will be
increased or decreased according to the legacy multiplier to ensure an
equivalent resultant load. For example, assume you have a nonlinear analysis
model (created in version 2012 or earlier), where a hydrostatic load with a fluid
density of 0.025 and a multiplier of 3 has been applied. When this model is
opened in version 2013 or newer, the specified fluid density will be 0.075 (3 x
0.025).

Refer to Figure 1 below. Point S represents any point along the top surface of the fluid. The
vector V is the direction of increasing fluid depth. The dark red lines are the surface of
elements forming a tank. The pressure acting on the wall (Pv) is a function of the depth,
increasing linearly as you move downward from the fluid surface. Since the default load
direction is normal to the surface of the elements, the pressure along the bottom of the tank
acts vertically.

Figure 1: Hydrostatic Pressure


In the Creating Hydrostatic Pressure Object dialog box, there is a Pressure Type option.
The choices are as follows:

Normal to the surface: The magnitude of the hydrostatic pressure =


(fluid density) x (depth below fluid surface), and the direction is normal to
the surface of each plate element. See Figure 2(a) below.

Full pressure in horizontal: The magnitude of the hydrostatic pressure


is calculated as usual, but applied in the horizontal plane. There is no force
component parallel to the vector defined by the Surface Normal of Fluid
vector. In other words, the direction is normal to the Surface Normal of
Fluid, regardless of the slope of the element surfaces. An example of when
this option might be used is to simulate the lateral soil pressure acting
against an inclined or curved retaining wall. See Figure 2(b) below.

Horizontal component only: This indicates that only the horizontal


component of the hydrostatic pressure is to be applied (vertical
component = 0). The magnitude of the pressure = (fluid density) x (depth
below fluid surface) x sin(angle between the surface normal direction of
the element and the Surface Normal of Fluid vector). The load direction is
normal to the Surface Normal of Fluid vector, regardless of the slope of the
element surfaces. See Figure 2(c) below.

Note: If the element surface is horizontal (such as along the bottom of a flatbottomed tank), then the normal direction for the element surface and the fluid
normal direction are parallel, so the horizontal component of the hydrostatic
pressure is zero. In this situation, the applied hydrostatic pressure is 0 for
Pressure Type selections of Full pressure in horizontal or Horizontal
component only.

(a) Normal to the surface

(b) Full pressure in


horizontal

(c) Horizontal component


only

Figure 2: Types of Hydrostatic Pressure

Visualization of Hydrostatic Pressure Loads:


For CAD-based models, arrows are displayed along surfaces to indicate applied hydrostatic
pressure loads in the FEA Editor. The arrows point in the surface normal or horizontal
direction (depending upon the load option you select). However, the lengths of the arrows are
not scaled to indicate the varying magnitude of the pressure as the fluid depth increases.
Arrows are not displayed for elevations above the fluid surface. Therefore, the hydrostatic
pressure arrows do not necessarily appear over the entirety of each selected surface.
For non-CAD-based models, H glyphs are displayed along the surfaces in the FEA Editor to
indicate hydrostatic pressure loads.
Hydrostatic pressures are converted to nodal forces during the solution phase. For all models,
the Results environment displays arrows of varying length, indicating the load direction and
the pressure variation at different fluid depths. You can check a model prior to solving it to
verify the proper hydrostatic loads.

Notes on General Shell Elements:


Nonlinear General Shell elements take the thickness of the element into account for pressure
loading.
(The other planar elements that support hydrostatic pressure loads plate, membrane, corotational shell, and thin shell consider the pressure to be applied at the midplane. The type
of shell element is set using the Element Formulation selector within the Advanced tab of
the Element Definition dialog box.)
As mentioned previously, general shell elements have options to apply the hydrostatic
pressure to the Top side, Bottom side, Both Sides, or Neither side.

Although the areas of the top and bottom sides of the element are equal in the stress-free
condition, large displacement effects can stretch the two surfaces differently. Thus, although
uniform pressures of -1000 on the top and 1000 on the bottom may appear to be identical
graphically, the results can be different. A similar situation occurs with hydrostatic loads. In
addition, for inclined or curved surfaces, taking the thickness into consideration changes the
effective fluid depth where the fluid contacts the elements, and therefore affects the
hydrostatic pressure (depending upon whether the load is applied to the top or bottom face of
the elements). See the figures below.

(a) Planar element, one with a negative pressure applied to the top side of the element (left
side of figure) and one with a positive pressure applied to the bottom side of the element
(right side of figure). In the stress-free condition, the area of the top side and bottom sides are
the same. (The element normal point is indicated by the X.)

(b) As the elements stretch, the area of the top and bottom sides also stretch. Thus, the total
force due the same pressure on the top as on the bottom may be different. In this example, the
top side stretches more than the bottom side, so the force in the model with the pressure on
the top is higher than the force in the model with the pressure on the bottom.

(c) The above image demonstrates how the thickness can affect the hydrostatic pressure for
general shell and plate elements. The X represents the element normal point. Notice that the
range of fluid depth along the bottom surface (Db1 through Db2) differs from the range of
fluid depth for the top surface (Dt1 through Dt2). Therefore, the hydrostatic pressure load
will differ depending upon whether it is applied to the top or bottom surface of the planar
element. This effect will hold true for general shell elements along inclined, curved, or
horizontal surfaces.

Tip: How to Combine Constant and Hydrostatic Pressures


A common occurrence is the combination of a constant pressure (P) and a hydrostatic
pressurefor example, a tank partially filled with water and pressurized with air above the
water. For 2013 and newer software versions, you can apply multiple surface loads to a single
surface, or group of surfaces, in a linear static stress analysis. Apply P as a surface pressure
load and add the hydrostatic load to the same surface or surfaces.
However, if you are setting up a nonlinear analysis, or if you anticipate the possibility of both
linear and nonlinear analyses (using separate design scenarios, for example), a different
method is required. Nonlinear analyses do not support multiple surface loads. There are two
methods of applying the combined constant pressure and hydrostatic pressure loads.
1. The first method is to specify a free surface point that is at a higher
elevation than the actual fluid surface. At the top of the fluid and above,
the pressure is P and the applicable equation is...

P = (coordinate of higher free surface - coordinate of actual fluid surface)*(fluid


density).
All values are known except for the coordinate of the higher free surface, so calculate
this value and enter it for the hydrostatic pressure. Rearranging the terms of the prior
equation, solve for the higher free surface as follows:
Coordinate of higher free surface = P / (fluid density) + (coordinate of actual fluid
surface)

constant pressure + hydrostatic pressure = hydrostatic pressure of greater depth


Naturally, the surface numbers of the model may need to be adjusted so that the
hydrostatic pressure is only applied where needed (below the water level) and not in
the dotted region of the figure (above the water level). A constant pressure is applied
to the surface above the water level. For CAD-based models, splitting the surfaces

within the CAD application is the preferred method of doing this, but the surface
attributes of the lines can also be changed in Autodesk Simulation.
2. The second solution to this problem is to apply a Surface Variable Load.
Define an equation, as a function of the appropriate coordinate direction,
that will produce the desired linearly increasing pressure. As for the
preceding example, the surface of the CAD or FEA model may have to be
split to contain the load to the desired region. For more information
regarding surface variable loads, refer to the Variable Pressures page.
Loads from File

Sometimes, you have loads from a source other than the software or from the results of a
prior analysis that you want to apply to the model. You can use Loads from File to do so.
Loads from File transfers loads from a text file or from a prior analysis results file to the
model. The command facilitates various multiphysics scenarios and is also useful when you
find it easier to calculate the load in a spreadsheet or other in-house program. In such cases,
import the load rather than entering it one-by-one through the interface. For example, you
used an in-house program to calculate the temperature distribution of a model due to laser
light passing through the components, and you want to apply those temperature results to a
stress analysis model.
Tip: You can also import reaction forces from an electrostatic model, or
temperatures from a heat transfer model, using the Loads from File dialog box.

The procedure to apply a load from another file, whether an Autodesk Simulation results file
or a text file, is as follows:
1. Select Setup
Loads
Loads from File from the ribbon, or, with
nothing selected, right-click in the display area of the FEA Editor and
choose the Loads from File command.
2. Press the browse button in the Results File column. A dialog appears
which allows you to select the results file. Use the Files of type: pulldown to select the appropriate type of file to read for the loads. See the
details below for each file type.
3. Select the file and click the Open button.
4. Only the loads from a single load case in the selected file can be applied to
a specific index (row of the spreadsheet). Select the load case in the Load
Case column. In some cases, it is appropriate to import multiple sets of
loads into a single model by pressing the Add Row button. The loads
assigned to the other rows of the spreadsheet can be from different load
cases (or time steps) in a single file, or the loads can be from different
files.
5. The loads will be placed in a specific load case or load curve in the
analysis. Enter the load case/load curve in the Structural Load Case
field. (Note: the name of the column may change with the analysis type.)

6. For the loads to be multiplied by a constant value before being applied to


the model, specify the constant value in the Multiplier column.
7. Click OK.

The imported loads do not appear in the FEA Editor. They appear in the Results environment
after doing a Check Model or performing the analysis.
Note:

Some loads, such as nodal forces, support multiple loads at the same
node. For example, applying one nodal force of [0,100,0] and another
nodal force of [0, 50,50] will result in 150 in the Y direction and 50 in the Z
direction.

Other loads, such as initial nodal temperatures, support only one load at
the same node. If multiple initial temperatures are loaded on the same
node, the processor will use only one of the temperatures.

Unless indicated otherwise, the results in the selected file are converted from the Model Units
of the results model to the Model Units of the current model. If the Model Units of the results
model cannot be determined, then it is assumed that no conversion is required.

Importing an Electrostatic Reaction Forces File:


The reaction forces from an electrostatic analysis can be imported into a stress analysis using
the Loads from File command. Refer to the Force page for details on this type of load
transfer.

Importing Temperature Results File:


The temperatures from a steady state or transient heat transfer analysis can be imported into
various stress analyses or transient heat transfer analyses using the Loads from File
command. Refer to the Temperature page for details on transferring temperature results to a
structural analysis. For more information regarding transferring temperatures to a transient
heat transfer analysis, see the Analyses Parameters: Transient Heat Transfer page.
Note: There are two methods that can be used to transfer temperatures into a
structural or thermal analysisone via the Analysis Parameters dialog box and
one via the Loads from File command. Both methods are discussed within the
above-referenced pages.

Importing ASCII XYZ Results File:


Selecting a file type of ASCII XYZ Results File will import nodal-based loads specified by
X,Y,Z coordinate and apply the loads to the nodes in the model with the same coordinate. The
tolerance used for the comparison is 1E-6. (If more than one node has the same coordinate,
the load is applied to the first node.) The file to import must have an extension of .xyz to be
recognized.
Any type of nodal load supported can be defined in the ASCII file.

The format of this file must be as follows. Other than the addition of the X, Y, Z coordinate of
the load, the format of the file must be identical to the nodal condition table (nodecond.dbf
located in the modelname.ds_data\design scenario\ds.mod folder). Each item needs to be on
the same row of the file and separated by a comma.

Nodal condition Type

X coordinate of the load

Y coordinate of the load

Z coordinate of the load

X value of the load

Y value of the load

Z value of the load

Load case number for the load. Remember: only load case numbers in the
file that match the load case number selected in the user interface are
transferred.

Property Identification of the load. This is an ID number which points to


AnalysisTypeMaster.Loading.Prop(Prop_ID) which enables the user to add
an arbitrary number of additional properties to this load. Normally this
value is 0.

Note: The result or value of the load in the ASCII file must be in the same units as
the current Model Units. The values will not be converted when imported.

Importing ASCII Nodal Condition Data File:


Selecting a file type of ASCII Nodal Condition Data File will import nodal-based loads and
apply them to the specified node. The file to import must have an extension of .nod to be
recognized.
Any type of nodal load supported can be defined in the ASCII file. The format of this file
must be identical to the nodal condition table (nodecond.dbf located in the
modelname.ds_data\designscenario\ds.mod folder).
Note: The value of the load in the ASCII file must be in the same units as the
current Model Units. The values will not be converted when imported.

Importing ASCII Element Condition Data File:


Selecting a file type of ASCII Element Condition Data File will import element-based loads
and apply them to the specified element. The file to import must have an extension of .ele to
be recognized.

Any type of element load supported can be defined in the ASCII file. The format of this file
must be identical to the element condition table (elemcond.dbf located in the
modelname.ds_data\designscenario\ds.mod folder).
Note:

The value of the load in the ASCII file must be in the same units as the
current Model Units. The values will not be converted when imported.

When importing the loads from an ASCII file, additional input may be
required in the model to utilize the loads. For example:
o

Certain types of loads, such as pressure (stress analysis) and


convection (thermal analysis) use a global multiplier to activate the
load. Be sure that the global multiplier is set on the Analysis
Parameters screen.

Don't forget to define the number of load cases or to define the load
curves as necessary for the imported loads. For example, if the
linear stress model is set up for 2 load cases and a load imported
from an ASCII file is assigned to load case 5, the load on load case 5
will not exist: the model has only 2 load cases! To create the
additional load cases in this example, add three rows on the
Multipliers tab of the Analysis Parameters screen.

Some loads may require input in the model table (model.dbf) or


other tables to fully define the load.

Nodal Weights

A Nodal Weight (or lumped mass) is a load that can be applied to a node in most linear and
nonlinear structural analyses. It is used to represent the mass of an attached part or
subassembly that is not included in the actual model geometry.
A Nodal Weight can be used to resist the translation or rotation of a node. In order to resist
rotation, the Nodal Weight must be attached to an element that supports rotational degrees of
freedom (such as beams, plates, and shells). For elements with five DOF, such as plates
(which lack the rotational DOF about the axis normal to the element), a Nodal Weight will
only effectively provide rotation resistance about the other two axes.

Apply Nodal Weights


If you have nodes selected, you can right-click in the display area and select the Add pull-out
menu. Select the Nodal Weight command. You can also access this load via the ribbon
command, Setup Loads Weight.
Select the appropriate radio button in the Mass Input section to determine if the Nodal
Weight input values are defined in units of force or mass (mass = weight/gravity). For
nonlinear analyses, only mass units are permitted (the Units of force selection is grayed-out).

If the Nodal Weight is equally effective in all translational directions, activate the Uniform
check box and specify the magnitude of the mass in the X Direction field of the
Mass/Weight section. Typically, a mass will be assumed to act equally in all directions for
the majority of applications. If the mass/weight is assumed to have different magnitudes
along the three translational directions, deactivate the Uniform check box and specify the
appropriate values in the X Direction, Y Direction, and Z Direction fields in the
Mass/Weight section. For example, if a mass (like a car) is sitting on wheels on a ship deck,
you may want to assume that deck motion along the direction of travel does not effectively
accelerate the mass of the car in that direction (because the wheels rotate easily). In such
cases, enter zero or a reduced magnitude for that direction.
If the Nodal Weight is to be effective in rotational directions, specify the appropriate values in
the X Direction, Y Direction, and Z Direction fields in the Mass Moment of Inertia
section. Here, with the exception of spherical objects or regular cubes, the mass moment of
inertia will normally vary considerably about the three axes.
How the weight/mass and mass moment of inertia behave in the various linear analysis types,
for global or local coordinate systems, and when negative values are specified is summarized
in the following table:
Analysis Type

Static Stress
with Linear
Materials

Global
Coordinates

Local
Coordinates

Mass Weight

With gravity,
masses are
converted to
forces by Fi=m i x
g i , where i is the
X, Y, and Z
directions and g
is the gravity
constant times
the multiplier.

With centrifugal
loads, masses are
converted to
forces by F i =mi x
ai, where i is the
X, Y, and Z
directions and a
is the
acceleration (r x
w 2 ).

Mass behaves as if
input is in global

Mass Moment of
Inertia
With centrifugal
acceleration, inertias are
converted to torques by
T i =I i x i , where i is the
X, Y, and Z directions
and is the angular
acceleration.

With centrifugal
acceleration, inertias are

Analysis Type

Mass Weight

coordinates, not local


coordinates.

Mass Moment of
Inertia
converted to torques by
T i =I i x i , where i is the
appropriate direction
and is the angular
acceleration.

Negative
Negative mass/weight and mass moment of inertia
Mass or
input values are converted to positive values
Weight Input during the solution. Therefore, results are identical
for positive and negative values. A warning
message appears in the analysis log for each
value converted.
Linear
Natural
Frequency
(Modal)

Global
Coordinates

Masses follow global


coordinate system and
effect the vibration in
the corresponding
direction.

Inertias follow global


coordinate system and
effect the vibration in
the corresponding
direction.

Local
Coordinates

Masses follow local


coordinate system and
effect the vibration in
the corresponding
direction.

Inertias follow local


coordinate system and
effect the vibration in
the corresponding
direction.

Negative
Negative mass/weight and mass moment of inertia
Mass or
input values are converted to positive values
Weight Input during the solution. Therefore, results are identical
for positive and negative values. A warning
message appears in the analysis log for each
value converted.
Linear
Natural
Frequency
(Modal) with
Load
Stiffening

Global
Coordinates

Load stiffening
effects due to the
lumped mass are
not accounted
for.

Masses follow
global coordinate
system and effect
the vibration in
the
corresponding

Inertias follow global


coordinate system and
effect the vibration in
the corresponding
direction.

Analysis Type

Mass Weight

Mass Moment of
Inertia

direction.
Local
Coordinates

Local coordinate
systems not supported.

Local coordinate systems


not supported.

Negative
Negative mass/weight and mass moment of inertia input
Mass or
values are interpreted as positive FORCE values of the
Weight Input same magnitude, regardless of the Mass/Force Units

selection. No warning messages are produced in the


analysis summary or log files. In other words, if you
specify a mass of -500, the result will be the same as for a
positive force of 500. The resultant nodal load will differ
by a factor of g.
Recommendation: To avoid confusion and potential
model setup errors, do not enter negative input values.
Critical
Buckling
Load

Nodal Weights are not supported for this analysis type (neither
masses/weights nor mass moments of inertia).

Transient
Global
Stress (Direct Coordinates
Integration)

Inertial effects follow


the global coordinate
system.

Inertias follow global


coordinate system and
effect the motion in the
corresponding direction.

Local
Coordinates

Inertial effects follow


the local coordinate
system.

Inertias follow local


coordinate system and
effect the motion in the
corresponding direction.

Negative
Negative mass/weight and mass moment of inertia
Mass or
input values are converted to positive values
Weight Input during the solution. Therefore, results are identical
for positive and negative values. A warning
message appears in the analysis log for each
value converted.
Nonlinear
Analyses

Global
Coordinates

Inertial effects follow


the global coordinate
system.

Inertias follow global


coordinate system and
effect the motion in the
corresponding direction.

Analysis Type

Local
Coordinates

Mass Weight

Inertial effects follow


the local coordinate
system.

Negative
Negative
Mass or
masses/weights
Weight Input produce reaction forces
1
in the opposite direction
relative to positive input
values. For example, a
force in the direction of
the gravity vector acts
on a positive nodal
mass when gravity is
applied to the model.
For negative values, the
force will act in the
opposite direction from
the gravity vector.

Mass Moment of
Inertia
Inertias follow local
coordinate system and
effect the motion in the
corresponding direction.
Negative mass moments
of inertia produce
reaction moments acting
in the opposite direction
relative to positive input
values. For example, a
positive inertia produces
a torque that opposes
the rotational motion of
the mass. For a negative
value, the torque acts in
the same direction,
assisting the rotational
motion.

Note: While negative masses, weights, or mass moments of inertia have theoretical
significance, and their effects can be quantified in a nonlinear analysis, there really are no
examples of this behavior in nature. Therefore, you will only need to use positive nodal
weights when modeling real-world phenomena.
Note: See the comments under the Application of Loads and Constraints at
Duplicate Vertices heading on the Loads and Constraints page for information
about how nodal loads are applied at duplicate vertices.

Comments Regarding Linear Dynamics Restart Analyses:


Frequency Response, Random Vibration, Response Spectrum, Transient Stress (Modal
Superposition), and DDAM analyses are all based on a prerequisite modal analysis. The unscaled response of the structure (vibration mode shape results) are scaled according to the
specified excitation (ground motion) and other applicable loads. Even though nodal weights
cannot be applied in these analysis types, their effects can be included in the initial modal
analysis. Therefore, nodal weights affect the restart analysis results, because the natural
frequency results on which they are based are affected by the nodal weights.
Note that load stiffening effects due to gravity are not calculated for nodal weights in a
Natural Frequency (Modal) with Load Stiffening analysis. Therefore, you will have to apply a
nodal force in the direction of gravity in addition to each nodal weight to account for this
effect. Be sure to run a Natural Frequency (Modal) with Load Stiffening analysis first, if you
want load stiffening effects to be represented in your restart analysis results.

Moments

A moment is a load that is applied to a node or surface. A moment can be applied about any
direction specified by a vector.

Apply Moments
If you have nodes or surfaces selected, you can right-click in the display area,
select Add and click Nodal Moment or Surface Moment, as appropriate. You
can also apply moments via the ribbon - select a node or surface and click Setup
Loads
Moment.
Note: You can use nodal moments only on plate and beam elements, while you
can use surface moments only on plate, brick, and tetrahedron elements. Plates
do not support nodal moments about the axis normal to the face of the element.

Specify the magnitude of the moment that is to be applied to each selected object in the
Magnitude field and the direction of the moment in the Direction section. Click the Flip
Direction button (
) to invert the sign of the applied moment, reversing its direction. The
data you enter for nodal moments applies to each object. For example, if you select 10 nodes
and specify a 10 in-lb nodal moment about the X axis, you are applying 100 in-lbs (10*10) to
your model.
Specify the load case or load curve in which you want the moment placed in the Load
Case/Load Curve field. If you want a moment to be applied in multiple load cases, you can
copy it to a new load set and change the Load Case/ Load Curve value.
Tip: See the comments under the Application of Loads and Constraints at
Duplicate Vertices heading on the Loads and Constraints page for information
about how nodal loads are applied at duplicate vertices.
Pressures or Tractions

You can apply pressure or traction loads to CAD-based, hand-built, or 2D Mesh Generation
models. For linear structural analyses, pressures or tractions can be applied to surfaces of 2D,
plate, membrane, thin composite, thick composite, brick, and tetrahedral elements. For
nonlinear structural analyses, pressures or tractions can be applied to surfaces of pipe, 2D, 2D
kinematic, 2D hydrodynamic, shell, membrane, brick, tetrahedral, 3D kinematic, and 3D
hydrodynamic elements.
What does a pressure/traction load do?

Applies an even distribution load (force per unit area) over the selected
area.

A pressure can be applied to the linear and nonlinear structural elements


that are listed at the top of this page. The pressure will be applied normal
to the faces of brick, tetrahedral, plate, thin composite, thick composite,
shell, membrane, 3D kinematic, and 3D hydrodynamic elements. It will be
applied normal to the edges of 2D, 2D kinematic, and 2D hydrodynamic
elements. It will be applied as internal pressure to pipe elements.

For linear analyses, a traction can be applied to plate, brick, and


tetrahedral elements. A traction applies a pressure that is oriented in a
specific direction.

For nonlinear analyses, a traction can be applied to all of the nonlinear


structural elements that support pressures, except for pipe elements. The
traction will be a pressure oriented in a specific direction.

Note: Local coordinate systems are not supported for surface pressure or traction
loads. The direction radio buttons and vector components are solely based on
the global coordinate system.

Apply Pressures/Tractions
If you have surfaces selected, you can right-click in the display area and select the Add pullout menu and then choose the Surface Pressure/Traction command. You can also access this
command via the ribbon (Setup Loads Pressure). You can click the ribbon command
either before or after selecting the model surfaces where you want the load applied.
If you are performing a nonlinear analysis or a transient stress (direct integration) analysis,
select the load curve that the pressure or traction will follow in the Load Curve field. Press
the Curve button to define a load curve in the Load Curve Editor, or use the Setup Model
Setup Parameters dialog box.
For nonlinear analyses only:

Specify a value other than 1 in the Multiplier field to have an additional


multiplier applied to this load.

For the load to maintain the same relative orientation (with respect to the
model) as it deforms, activate the Follows Displacement check box.

Note: For CAD-based solid or surface models, pressure and traction loads are
visually represented in the FEA Editor using arrows that indicate the load
direction. For all other models, "P" glyphs are displayed at each node along the
loaded surfaces to indicate pressure or traction loads.

Pressures
To apply a normal pressure, select the Pressure button (this is the default). Specify the
magnitude of the pressure in the Magnitude field. Click the Flip Direction button (
invert the sign of the applied load, reversing its direction.

) to

For 2D, and solid elements, a positive pressure is directed into the element and a negative
pressure is directed away from the element.
For plate, thin composite, thick composite, and shell elements, a positive pressure points
away from the element normal point and towards the elements. A negative pressure points
towards the element normal point and away from the elements. The element normal point is
defined in the Orientation tab of the Element Definition dialog box.

Tractions
To apply a traction load to any of the supported elements, select the Traction button (it will
be grayed-out for non-supported elements). Specify the component of the traction in each of
the global directions in the X Magnitude, Y Magnitude, and Z Magnitude fields.

Notes on General Shell Elements:


Nonlinear General Shell elements take the thickness of the element into account for pressure
loading.
(The other planar elements that support hydrostatic pressure loads plate, membrane, corotational shell, and thin shell consider the pressure to be applied at the midplane. The type
of shell element is set using the Element Formulation selector within the Advanced tab of
the Element Definition dialog box.)
General shell elements have options to apply the pressure or traction load to the Top side,
Bottom side, Both Sides, or Neither side. The bottom side of the element is the side facing
the element normal point, as defined in the Element Definition dialog. A positive pressure
points into the element regardless of the side to which it is applied.
Although the areas of the top and bottom sides of the element are equal in the stress-free
condition, large displacement effects can stretch the two surfaces differently. Thus, although
uniform pressures of -1000 on the top and 1000 on the bottom may appear to be identical
graphically, the results can be different. A similar situation occurs with hydrostatic pressure
loads. In addition, for inclined or curved surfaces, taking the thickness into consideration
changes the effective fluid depth where the fluid contacts the elements, and therefore affects
the hydrostatic pressure (depending upon whether the load is applied to the top or bottom
face of the elements). See the figures below.

(a) Planar element, one with a negative pressure applied to the top side of the element (left
side of figure) and one with a positive pressure applied to the bottom side of the element
(right side of figure). In the stress-free condition, the area of the top side and bottom sides are
the same. (The element normal point is indicated by the X.)

(b) As the elements stretch, the area of the top and bottom sides also stretch. Thus, the total
force due the same pressure on the top as on the bottom may be different. In this example, the
top side stretches more than the bottom side, so the force in the model with the pressure on
the top is higher than the force in the model with the pressure on the bottom.
Remote Forces
The information in this section applies to linear and nonlinear structural
analyses. Specifically, Remote Force loads are supported in the
following analysis types:

Linear:

Static Stress with Linear Material Models

Natural Frequency (Modal) with Load Stiffening

Transient Stress (Direct Integration)

Critical Buckling

Nonlinear:

Mechanical Event Simulation (MES) with Nonlinear Material Models

Static Stress with Nonlinear Material Models

MES Riks Analysis

What does a Remote Force do?

Simulates the effects of a force applied at a point in space that is not on


the model.

Remote Forces are only applicable to model surfaces.

The nodal reaction forces that would occur at the model due to the remote
force are calculated, and nodal forces are applied to all of the nodes along
the selected surface or surfaces.

Remote Forces differ from Remote Loads and Constraints, in which


automatically-generated line elements connect the model to a point in
space, and a single nodal load or constraint is applied at the remote point .

Important: When a Remote Force is applied to multiple surfaces, the specified


load is NOT distributed over all of the selected surfaces, with each surface
receiving only a portion of the load magnitude. Rather, the full specified force is
distributed over each of the individual selected surfaces. For example, if a 200N
force is applied to three surfaces, the total applied load is 600N (200N per
surface times 3 surfaces). If you need a single load magnitude at a point in space
to be distributed over multiple surfaces, use the Remote Loads and Constraints
command instead of the Remote Force command.

Apply a Remote Force


1. Select one or more surfaces where the remote force will act on the model.
For example, choose the mounting surfaces for a motor bracket that is not
included in the model. Remote forces can be used to apply the effects of
the motor and bracket weight, plus any operating loads, at the selected
mounting surfaces.
2. Click the Setup
Loads
Remote Force ribbon command. Or, rightclick in the display area and choose Add
Surface Remote Force from
the context menu.
3. Specify the force Magnitude. Click the Flip Direction button (
the sign of the applied load, reversing its direction.

) to invert

4. Specify the location of the remote point by one of the following two
methods...
o

Type the coordinates in the X, Y, and Z input fields, or

Click the Point Selector button and then pick a vertex on the
model. This method works if a construction vertex has been added
at the remote force location. It also is useful if the desired point is at
a known offset from an existing model vertex. In the latter case,
adjust one or two of the coordinates to reflect the location offset
after choosing the reference vertex.

5. Specify the direction of the load by one of the following three methods...
o

Activate an X, Y, or Z radio button to specify a global axis direction.

Activate the Custom radio button and enter the components of a


unit direction vector in the X, Y, and Z input fields.

Click the Vector Selector button and then click two points on the
model in succession to indicate the desired vector direction.

6. Specify the Load Case / Load Curve number. Optionally, click the Curve
button to define a load curve for nonlinear analyses.

7. Click OK to apply the force and close the Remote Force Object dialog box.
Remote Loads and Constraints

What Do Remote Loads and Constraints Do?

Adds a nodal load or boundary condition to a point in space; a point not on


the model.

The point in space is connected to selected nodes on the model with line
elements.

You define the properties of the line elements as beam, truss, or similar
line elements.

The remote load or boundary condition is transmitted through the line


elements to the model.

Remote Loads and Constraints differ from Remote Forces, in which the
effects of a remote force are applied directly to the model surface using
automatically-calculated nodal forces (no line elements are created).

Since the Remote Load command generates new geometry and a node at the point in space,
you can add any number of additional objects at the new point.

Apply Remote Loads or Constraints


1. Use any of the selection commands (Selection
Select) to select where
the remote load or constraint is to be distributed onto the model. For
example, to apply a torque to the end of a shaft, select the vertices or
surface on the end of the shaft. Regardless of what is selected (part,
surface, edge, line), the vertices on the selection are used. Most
applications require the remote load or constraint to be distributed to
three or more vertices not in a straight line.
2. Right-click and choose the Create Remote Load & Constraint
command. This command is also available via the ribbon (Setup
Remote Load & Constraint).

Loads

3. Use either of the following methods to specify the location of the remote
load or constraint.
o

Type the X, Y, and Z coordinates of the desired Load Location. Or,

Select one vertex or construction vertex in the display area and


click the Use Selected Point button. The coordinates of the
selected vertex will be listed in the X, Y, and Z fields.

4. The part, surface, edge, or vertices over which the remote load or
constraint is to be distributed are already listed by virtue of starting the
command with them selected. However, to change the destination or
create a new remote load or constraint, do one or both of the following:

Select the new location or locations, and click the Add button.

Select one or more lines in the Load Destination list and click the
Remove button.

5. Specify the Part, Surface, and Layer attributes for the line elements that
will connect the remote load to the model. Generally, the default part
number is a new (previously unused) number or the same number as
previously generated remote load elements.
6. Click the Generate Elements button. The lines from the remote load
location to the model are created.
7. Click the Add Load button to see the list of nodal loads or boundary
conditions that can be added. Some of the available options may not be
suitable depending on the analysis type and the type of elements used for
the load elements.
8. If another remote load location is required, repeat the above steps starting
with step 3.
9. Click the Close button when finished applying all remote loads to all
remote locations.
Tip: While the Create Remote Load dialog box is opened, multiple loads or
boundary conditions can be added to the same node by using the Add Load
button more than once. After closing the dialog box, additional loads and
boundary conditions can be added to the same node by selecting the vertex at
the remote load location ( Selection
Select
Vertices), right-clicking, and
choosing the appropriate entry under the Add fly-out menu.

Define Remote Load/Constraint Line Elements


After the remote load or constraint line elements are created, use the browser (tree view) to
define the Element Type, Element Definition, and Material. Any line element type can be
chosen (beam, truss, gap, and so on), provided it suits the requirements of the analysis. Here
are a few guidelines to keep in mind:

Moments can be applied as a remote load, but they can only be


transmitted through beam elements. Truss elements, gap elements, and
other line element types that do not have rotational degrees of freedom
cannot transmit moments and torques. the joints of these element types
are pinned (no translation, rotation allowed). (See Getting Started:
Introduction to Autodesk Simulation FEA: Nodes and Elements for
additional information on transmitting loads, restraints, and degrees of
freedom.)

Imagine the array of load elements as being supported by boundary


conditions instead of connected to the model. The support reactions at
these hypothetical boundary conditions are the loads that are transmitted
to the model. The total of these support reactions equals the applied
loads, but the distribution of the forces and moments may be affected by
the stiffness of the load elements.

Stiff beam elements acts like a rigid structure attached to the model, so
the surface (or nodes) of the model where the remote load is distributed
tend to keep the same shape but move as a rigid surface. Weak beam
elements and truss elements transmits the load but do not completely
prevent deformation of the shape of the surface.

Example Use of Remote Loads and Remote Constraints


Figures 1 and 2 illustrates how to use remote loads to analyze a shaft (made from brick
elements) that is part of a gear train. In the FEA Model, the two boundary conditions at the
bearings prevent the shaft from rigid body translations in all directions. The bearing on the
left is held radially (Ty and Tz) and axially (Tx, to contain thrust loads), and the mounting of
the bearing on the right constrains radial translation (Ty and Tz) but allows axial movement
(floating bearing). Assume the bearings are spherical, so that they do not restrain rotation in
any direction.
If the load at the pinion were to be modeled with a force, the shaft would still be free to rotate
about the axial direction. This would lead to an unstable model and potentially wrong results.
A remote boundary condition at the pinion tooth that restrains the model in the tangential
direction prevents axial rotation (Rx) and produces the reaction force needed to balance the
gear load.

Figure 1: Diagram of Shaft With Gears

1. Remote Constraint attached to surface of bearing journal with a notranslation constraint at center of bearing (Tx, Ty, and Tz).
2. Remote Load attached to gear mounting surface with force applied at gear
tooth location.
3. Remote Constraint attached to pinion mounting surface with boundary
condition in tangential direction (Ty).
4. Remote Constraint attached to surface of bearing journal with axial and
radial constraint boundary conditions (Ty and Tz).

Figure 2: Equivalent FEA Model Using Remote Loads


Tip: If working with a CAD model, the Mesh
CAD Additions
Joint command
can be used for the bearings to create universal-type joints. The result is the
same geometry as created by the remote load command.
Temperatures

Temperatures are also applicable to heat transfer analyses but are significantly different
within that context. For more information, please refer to the following pages:

Initial Temperature

Controlled Temperature

A temperature can be applied to nodes, surfaces, or parts in a linear or nonlinear structural


analysis model and to nodes or surfaces in an electrostatic analysis model.
What Does a Temperature Do?

A surface temperature applies nodal temperatures to each node on the


surface, and a part temperature applies nodal temperatures to each node
in the part.

In linear and nonlinear structural models, temperatures are used for


thermal stress analysis. Temperatures cause thermal expansion of any
materials with non-zero coefficients of thermal expansion.

The nodes to which a temperature is applied are kept at the value


specified in the Magnitude field. This temperature is only applied to the
selected nodes and does not conduct through the material. (A thermal
analysis is required to calculate heat conduction or heat flow via
convection or radiation.)

The stress caused by the temperature is calculated from the difference


between the nodal temperature and the Stress Free Reference
Temperature, which is specified within the Element Definition dialog
box. The effect of this temperature difference depends on the Thermal
Coefficient of Expansion defined in the Material Specification dialog
box.

For the manually-applied temperatures to be accounted for in a linear or


nonlinear structural analysis, you must select the Model file option in the
Source of temperature drop-down box in the Thermal tab of the
Analysis Parameters dialog box.

For a linear analysis, you must also assign a Thermal multiplier in the
Multipliers tab of the Analysis Parameters dialog box.

For a nonlinear analysis, select the load curve that will control the
magnitudes of the temperatures as a function of time in the Nodal
temperature load curve index field within the Thermal tab of the
Analysis Parameters dialog box. A load curve must have been
previously defined for its number to appear in this drop-down menu. The
load curve can be used to vary temperatures from a steady-state heat
transfer analysis over time. For example, you can gradually increase the
applied temperature from zero to the input temperature values by varying
the load curve multiplier from zero to 1.
Important: For nonlinear stress analyses using transient heat transfer
results, the temperature as a function of time in the stress analysis is the
same as the temperature versus time from the thermal analysis. The
duration of the thermal simulation event must be equal to or greater than
the duration of the stress analysis event. In this case, the load curve will
typically have a constant value of 1 (applied temperature equals input
temperature from the thermal analysis throughout the event). This is
discussed further in the Use Load Curves to Control Temperature versus
Time in Nonlinear Analyses section at the bottom of this page.

For linear, nonlinear, and electrostatic analyses, temperatures affect the


material properties when temperature-dependent material models are
specified within the Element Definition dialog box.

Apply Temperatures
If you have nodes, surfaces, or parts selected, you can right-click in the display area and
select the Add pull-out menu. Select the Nodal Temperatures or Surface Temperatures or
Part Temperatures command, respectively. You can also access this command via the ribbon
(Setup Loads Temperature).
Specify the magnitude of the temperature that is applied to each selected object in the
Magnitude field.
Note: If different temperatures are applied to the same node, the last
temperature is used for the node. For example, if parts 1 and 2 are bonded
together, and if part 1 is assigned a temperature of 100 degrees, then part 2 is
assigned a temperature of 75 degrees, the nodes in common will have an initial
temperature of 75 degrees.

Apply Constant Temperatures to Entire Models


If you simply want to determine the thermal stress in a model due to a uniform temperature
change, it is not necessary to add nodal temperatures to the entire model. Instead, right-click
the Thermal heading under the Analysis Type heading in the browser and choose Edit. Note
that this heading may be gray before a temperature has been defined, but the command is still
available. Perform the following two steps:

Choose the Loads from FEA Editor option from the Source of
Temperature drop-down menu.

Type the desired temperature value in the Default Temperature field.


Any node in the model that does not have a nodal temperature applied is
set to this value. Any applied temperature overrides the default value.

Apply Temperature Profiles from Thermal Analyses


In some cases, the temperature profiles for a model have already been calculated using either
a steady-state or transient heat transfer analysis. If the geometry of the structural model is
identical to the thermal model, the thermal results can be used for the temperature profile.
There are two methods to do this.
Method 1:
1. For a linear or nonlinear stress analysis, right-click the Thermal heading
under the Analysis Type heading in the browser and select the Edit
command. For a natural frequency (modal) with load stiffening or critical
buckling load analysis, right-click the Analysis Type heading, select the
Edit Analysis Parameters command and go to the Thermal tab.
2. Then, select the type of heat transfer analysis that was previously
performed on this model in the Source of Temperature drop-down
menu. The choices are as follows:
o

Another Design Scenario in loaded file.

Another Simulation Mechanical file.

Autodesk CFD file. Please refer to the Using Autodesk CFD


Temperature Results as Thermal Loads page for more information
concerning this option.

3. If using another Simulation Mechanical file or an Autodesk CFD model,


press the Browse button next to the Filename and Design Scenario
heading. Then, navigate to and select the desired thermal analysis model
file. Specify which design scenario of the thermal source model to use
from the drop-down menu below the filename field.
If you are using a different design scenario within the same model file,
specify which one to use from the Use temperature from design
scenario pull-down menu.
Note: This menu will only list design scenarios for which the analysis type
is either steady-state or transient heat transfer. If a heat transfer analysis
was completed but the analysis type later changed to a non-thermal type,
the design scenario will not appear in the Use temperature from
design scenario pull-down menu, even though the thermal results file is
still present in the folder.
4. If you are using the results from a transient heat transfer analysis as a
load in a linear static stress analysis, specify which thermal analysis time
step to use as the source of temperatures for the stress analysis. Do this
using the Which step to use pull-down menu. The default is to use the
last time step.
Note: Method 1 permits the meshes to be different between the thermal model
and the stress model. See the paragraph Requirements for Different Meshes
on the Multiphysics page for details.

Method 2:
1. With nothing selected, right-click in the display area of the FEA Editor.
Select the Loads from File command.
2. Press the Browse button in the Results File column. A dialog box
appears so you can select the results file. Use the Files of type: pulldown to select Thermal Results File (*.to, *.tto).
3. Select the file with the temperature results and click Open.
4. Only the temperatures from a single load case or time step in the selected
file can be applied to the stress model. Select the load case or time step in
the Load Case from File column.
5. The temperatures are placed in load case 1 regardless of what number is
entered in the Structural Load Case field.

6. If you want the temperatures to be multiplied by a constant value before


being applied to the model, specify the constant value in the Multiplier
column.
7. Press the OK button.
Attention: Unlike other types of loads which can be applied to the model using
the Loads from File method, only one temperature can be applied to a given
node. You cannot specify multiple result files (or the same file multiple times)
and have the temperatures added together at the same node. Nor can you apply
different thermal results files to the same nodes for different load cases. Only the
last temperature applied to a node will be retained. You can use multiple
temperature results files only if each file pertains to a different portion of the
stress model.
Note: Method 2 requires the meshes to be identical between the thermal model
and the stress model because it transfers the temperatures by coordinates in the
overall model (not on a per-part basis). See the paragraph Requirements for
Same Meshes on the Multiphysics page for details.

Use Temperatures to Model Initial Strain Conditions in a Linear


Analysis
The easiest way to model an initial strain in a part is to use a temperature. If you know the
amount of existing strain and the properties of the material, you can apply the correct
temperature difference using the following procedure.
1. Using the basic equation for expansion, we know that the strain is equal to
the product of the temperature difference and the coefficient of thermal
expansion, :

2. Solving this equation for the temperature difference we get:

Using this process we can now calculate the temperature difference that we have to apply to
the model to simulate a prestrain condition based on the known strain and the material
properties.

Use Load Curves to Control Temperature versus Time in Nonlinear


Analyses
All temperatures are multiplied by the assigned load curve multiplier where the load curve is
set with the Nodal temperature load curve index drop-down. For a uniform rod, the
expansion due to the temperatures would be...
L = L**(LCM*T - Tref)
where

L is the change in length

L is the length

is the coefficient of thermal expansion, which is based on the


temperature T, not LCM*T

LCM is the load curve multiplier

T is the temperature at each node, either from temperatures explicitly


assigned to the nodes, or the default nodal temperature, or the results
from a steady state analysis

Tref is the stress free reference temperature.

Since a transient heat transfer analysis presumably has an accurate history of the temperatures
versus time, it may be unusual to read these temperatures and use a load curve multiplier
other than 1. In some cases though, such as when the model begins at a temperature other
than Tref, the impact that occurs due to the application of the temperatures at time 0 can be
reduced by ramping up the load curve multiplier over a portion of the event. This will help
the solution to converge.
Where steady-state thermal results are used as the source of temperatures for a nonlinear
structural analysis, use the load curve to gradually increase the temperature from Tref to the
input temperature values.
Variable Pressures

What Does a Variable Pressure Load Do?

A variable pressure is a load that is applied over an area; a surface. A


variable pressure can be applied in any direction specified by a vector or
can be applied normal to the surface.

A function can be defined to control the magnitude of the load at different


locations on the surface.

The variable pressure is converted into equivalent forces and applied to


the element nodes.

Apply Variable Pressures


If you have surfaces selected, you can right-click in the display area and select the Add pullout menu. Select the Surface Variable Pressure command. You can also access this
command via the ribbon (Setup Loads Variable Pressure).
Variable pressures can be applied to any meshed surface: specifically on plates, shells
composites, membranes, bricks, tetrahedra, 3D kinematic, and 3D hydrodynamic elements.
They are not available for 2D elements.

Select the coordinate system that the variable pressure follows in the Coordinate system
drop-down menu. All local coordinate systems defined in the FEA Editor environment is
listed in addition to the global coordinate system. For example, if you want a surface variable
pressure to start at zero at an edge of a surface, you should create a coordinate system that has
the origin at one of the nodes along that edge. Select that coordinate system in this drop-down
menu.
Note: All occurrences of the function must use the same coordinate system. If
you change the coordinate system after the function is defined, then all surfaces
in the model that use the same function is also updated to use the newly chosen
coordinate system.

If the direction of the variable pressure will always be normal to the surface, select the
Normal to surface radio button in the Load orientation section. If the variable pressure is
applied in a constant direction, select the Traction radio button in the Load orientation
section and define the direction as a vector in the X, Y, and Z fields.
When creating a new load, the Active function drop-down menu is set to New function.
Either type a name to create a new function, or use the drop-down menu to select an existing
function. If you change the parameters of the function, the changes are applied to every
variable pressure in the model that uses that function.
Define the function in the Expression field. The equation represents a pressure applied to the
surface. Use the variables r, s, and t in the expression; note how the definition of these
variables changes depending on the coordinate system as shown on the dialog box. You can
use basic operators such as + - * / ( ) and ^. Click the Available Primitives >> button to
access several common functions.
Click the View button to see a graphical view of the load along a specific direction. You can
select the radio button for the variable plotted along the abscissa axis (the X axis) of the
graph. Define the other two coordinates if they are used in the expression. For example, if
you have a load that increases linearly in the X direction, select the X radio button. It
generates a straight line at a constant slope. Since the load does not vary in the Y or Z
directions, it does not matter what values are entered for these coordinates. If the expression
were 5*(X+Y) and you chose the X radio button to see how the pressure changes versus the
X coordinate, then you need to enter a specific value for Y. The graph will show how the
pressure varies in the X direction based on the entered Y coordinate. You can view the graph
at different Y coordinates by changing the coordinate and pressing the Recalculate button.
In the Load case / load curve field, specify the load case in which you want the variable
pressure placed (for linear analyses). If you are going to perform a transient stress (direct
integration) analysis or a nonlinear analysis, this field contains the number of the load curve
that will be used to control the variable pressure throughout the analysis.
Since the variable pressure is converted into nodal forces, the load in the Results environment
may not appear to follow the function exactly. If the mesh is not evenly spaced, numerous
forces of low magnitude may be applied in an area with a fine mesh and fewer forces of
higher magnitudes may be applied in an area with a coarse mesh. Also, nodes that are not

shared by multiple elements have a force with a lower magnitude than the others. The node
gets a force applied from each element that it is shared by.
Tip:

When modifying a surface variable pressure, it is suggested to use the


same Display Units as when the load was created. (Right-click the Display
Units in the Unit System branch of the tree view and choose Activate.)
Viewing the load in other units shows conversion factors since the
equation must give the same result regardless of the units.

For example, the equation: 0.0361 [lbf/in^3] * (12.5 [inch] - Z [inch])


represents a hydrostatic pressure in units of psi, increasing in the negative
Z direction, starting at an elevation of Z=12.5 inches. (The units are
shown in brackets [ ] in this example to make the discussion easier to
follow; the software does not show the units for each term.) When the
Display Units are set to SI and the load is modified, the equation is now
0.0361 [lbf/in^3] * (12.5 [inch] - 39.37 [inch/m] * Z [m]) * 6894.8
[Pa/psi]

A quick hand calculation shows that these two equations give identical
pressures in the respective units, but interpreting the second equation can
be more challenging.

Example of Variable Pressure


We will set up and view a model of a 5 inch cube with the following variable loading.

First we should set up a rectangular local coordinate that has the origin along the edge of 0
pressure and the local X axis parallel to the direction of increasing load. Next we select the
surface and right-click in the display area and select the Add Surface Variable Pressure
command. The variable pressure parameters should be set as shown below.

Click the View button to show the following graph along the X direction.

Since the model does not vary in the Y or Z directions, the values in the y and z coordinate
fields are not relevant. For verification purposes, the value half way along the surface should
be constant at -30*(2.5)^2 = -187.5. Select the T radio button and type 2.5 in the R field.
Press the Recalculate button. The following graph should appear.

If we view the model in the Results environment, it appears as shown in the following image.
The reason for the shorter arrows along the edge is because there are fewer elements sharing
those nodes.

Voltages
Voltages are also applicable to electrostatic analyses but are significantly
different within that context. Please refer to the Applied Voltage page for more
information.

A voltage can be applied to nodes or surfaces of a model.


What Does a Voltage Do?

A surface voltage applies nodal voltages to each node on the surface.

A voltage is used for a voltage induced stress analysis. The node that a
voltage is applied to is kept at the value specified in the Magnitude field.
This voltage only applies to that node. It does not conduct through the
material. Either a Piezoelectric or General Piezoelectric material
model must be selected in the Material Model drop-down menu of the
Element Definition dialog box for nodal voltages to affect a structural
analysis.

The stress caused by the nodal voltage is determined by the values


entered in the Piezoelectric tab of the Element Material Specification
dialog box.

For the manually-applied voltages to be accounted for in the analysis, the


Model file option must be selected in the Source of voltages drop-down
box in the Electrical tab of the Analysis Parameters dialog box.

For a linear analysis, you must also assign a Electrical multiplier in the
Multipliers tab of the Analysis Parameters dialog box. This value is
multiplied by the Magnitude and the product is applied to the node.

For a nonlinear analysis, select the load curve that will control the
magnitudes of the voltages as a function of time in the Nodal voltage
load curve index field within the Electrical tab of the Analysis
Parameters dialog box.

Apply Voltages
If you have nodes or surfaces selected, you can right-click in the display area and select the
Add pull-out menu. Select the Nodal Voltages or Surface Voltages command. This
command is also available from the ribbon (Setup Loads Voltage).
Specify the magnitude of the voltage that is applied to each selected object in the Magnitude
field.
Note: See the comments under the Application of Loads and Constraints at
Duplicate Vertices heading on the Loads and Constraints page for information
about how nodal loads are applied at duplicate vertices.

Apply Constant Voltage to Entire Models


If you simply want to determine the stress in a model due to a uniform voltage change, it is
not necessary to add nodal voltages to the entire model. Instead, right-click the Electrical
heading under the Analysis Type heading in the browser. Type the value in the Default nodal
voltage field. Any node in the model that does not have a nodal voltage is set to this value.

Apply Voltage Profiles from Electrostatic Analyses


In some cases, the voltage profile for a model has already been calculated using an
electrostatic analysis. If the geometry of the structural model is identical to the electrostatic
model, the voltage results can be used for the voltage profile.
Right-click the Electrical heading under the Analysis Type heading in the browser. Select
the Electrostatic analysis option in the Source of nodal voltages drop-down Menu. Click
the Browse button next to the Voltage data in file field and navigate to the electrostatic
results file.
Note: This method permits the meshes to be different between the electrostatic
model and the stress model. See the paragraph Requirements for Different
Meshes on the Multiphysics page for details.
General Constraints - Boundary Conditions

A boundary condition can be applied to nodes, edges or surfaces of a model.


What Does a Boundary Condition Do?

A boundary condition holds a node and prevents it from translating and/or


rotating. It acts like a rigid support between the model and the ground.

An edge or surface boundary condition applies nodal boundary conditions


to each node on the edge or surface.

When the model is being analyzed, an equation is generated for each


degree of freedom of each node. If a boundary condition is applied to a
node no equation is generated for that node because it experiences no
translation or rotation.

To model a cantilever beam, constrain both the translation and rotation at


the fixed end.

To model a simply supported beam (using beam elements), constrain the


three translations and axial rotation at one end and constrain the two
transverse translations at the opposite end. (If the beam were parallel to
the X axis, the constraints are TxTyTzRx at one end and TyTz at the other.
The Rx constraint is necessary to provide torsional stability. Freeing Tx at
one end is optional for a linear static analysis, since this analysis type
does not account for large displacement effects. So, the two supports do
not move closer together regardless of the amount of deflection.)

Each element type supports certain degrees of freedom (DOF). If you


apply a boundary condition to a DOF on an element that does not support
that particular DOF, the boundary condition is ignored. For example, truss
elements cannot resist rotation because they have no rotational DOF. They
behave as if both ends are ball-joint connections. If you place a fixed
boundary condition on an end of a truss element, the three rotational
constraints are ignored. Similarly, brick and tetrahedral elements have
only translational DOF. Rotational constraints are ignored for these solid
elements too.

Apply Boundary Conditions


If you have nodes, edges or surfaces selected, you can right-click in the display area and
select the Add pull-out menu. Select the Nodal Boundary Condition, Edge Boundary
Condition or Surface Boundary Condition command. This command can also be accessed
via the ribbon (Setup Constraints General Constraints). Edge boundary conditions
can only be applied to parts that originated from CAD solid models or the 2D Mesh
Generation.
Either press one of the buttons in the Predefined section or activate the appropriate check
boxes in the Constrained DOFs section. The Fixed button activates all six check boxes. The
Free button deactivates all six check boxes. The No Translation button activates the Tx, Ty,
and Tz check boxes. The No Rotation button activates the Rx, Ry, and Rz check boxes.
The remaining six buttons apply symmetric or antisymmetric boundary conditions. See the
Model Symmetry page for more information.
Tip: In nonlinear analyses, prescribed displacements can be used in place of
boundary conditions when the node needs to be released at some time. Use the
death time on the Prescribed displacement to specify when the node will be
released. See the Prescribed Displacement page for details.
Note: See the comments under the Application of Loads and Constraints at
Duplicate Vertices heading on the Loads and Constraints page for information
about how nodal loads are applied at duplicate vertices.

Multi-Point Constraints

Multi-point constraints (MPCs) are an advanced feature where you connect different nodes
and degrees of freedom together in the analysis. They are often used to simulate a boundary
condition effect when regular boundary conditions do not provide the correct behavior.
One use of MPCs is a master and slave situation: the displacement at node X (the slave node)
is needed to be the same as at node Y (the master node). In Figure 1, a portion of a long
vessel is modeled. The left side uses symmetry boundary conditions; this restrains the model
in the Z translation and simulates the portion of the tank to the left. For a long vessel, the
portion of the tank to the right side of the model forces those nodes to remain in a plane. A
symmetry boundary condition would work except that these prevent axial growth or
contraction in the tank. Instead, MPCs are used to indicate that the Z displacement of all the
nodes are equal (but not necessarily 0). Similarly, temperatures in a thermal analysis and
voltages in an electrostatic analysis can be the basis of MPCs.
Detail:

Z symmetry
conditions
restrain the
left face.
Those nodes
do not move
in the Z
direction.

MPC
conditions
restrain the
right face to
remain in a
plane; the
nodes move
together in the
Z direction.

Without MPC or
other boundary
conditions, the
nodes on the
right face are
free to deflect
due to the
loads. This does
not accurately
simulate the
portion of the
vessel not
included in the
analysis.

Figure 1: Use of Multi-point Constraint


Note: Among the Linear analysis types, multi-point constraints are available only
for the Static Stress with Linear Material Models, Natural Frequency (Modal), and
Transient Stress (Direct Integration). Also, the linear analysis types that use
modal results (such as Response Spectrum, Random Vibration, and so on)

produce results that include the effects of the MPCs. For thermal analyses, MPCs
are available for both Steady-State and Transient Heat Transfer. Likewise, both
types of electrostatic analyses, Current and Voltage as well as Field Strength and
Voltage support MPCs.

The input for multi-point constraint is an equation with the following format:

where

i is the ith term of the equation

M i is a multiplier for term i of the equation

DOF@Node i is a degree of freedom (DOF) at a specific node for term i. The


type of DOF depends upon the analysis type (translational or rotational
displacements for linear structural analyses, temperature for thermal
analyses, and voltage for electrostatic analyses).

n is the number of terms in the equation

Const is the constant that the equation equals. This value is often zero.

If the equation involves any units, they are written using the Model Units. The MPC
equations do not use the Display Units.

Enter the MPC Equation


1. In the FEA Editor, write down the vertex numbers and associated DOF(s)
required for the MPC equations. To get the vertex numbers, use the
Selection
Select
Vertices to select one vertex, then right-click and
choose Inquire. (Or, just holding the mouse pointer over a vertex with
show the properties in a tool tip.)
2. For all supporting analysis types, the Multi-Point Constraint command is
found in the Setup tab of the ribbon.
o

For linear structural analyses, the command is in the pull-out


portion of the Constraints panel.

For thermal analyses, the command is in the pull-out portion of the


Thermal Loads panel.

For electrostatic analyses, the command is in the pull-out portion of


the Loads panel.

Also, with nothing selected on the model, you can right-click in the display
area and choose Add
Multi-Point Constraint. Use the Define MultiPoint Constraints dialog box to enter all the terms of the previous
equation.
3. Click the Add button to create a new constraint equation; the name of the
equation is automatically filled in. Or use the Equation name drop-down
menu to select an existing equation to edit.
4. Specify the constant of the equation in the Constant field. (Const in the
above equation.)
5. Use the Add Row button to add as many rows to the spreadsheet as
required to specify each term of the equation. In each row, enter the
Multiplier and Vertex ID ( the vertex number written down from step 2)
and the appropriate DOF (degree of freedom). Use the available dropdown list for linear analyses. The DOF is fixed for thermal and electrostatic
analyses (DOF = Temperature and Voltage, respectively).
Note: If a vertex is assigned to a local coordinate system, the DOF
selected is also in the local coordinate system. For example, X translation
refers to the radial translation direction in a cylindrical coordinate system,
Y translation refers to the tangential direction, and so on.
6. Choose the Solution method and set the Penalty multiplier (the
Penalty multiplier is used for the Penalty method). You can select from the
following Solution methods:
o

Automatic

Penalty Method

Condensation Method

Note: You can find recommended values for the Penalty multiplier in the
respective Set Up topics. Search on "Penalty multiplier" for quick access to
the topics.
Note: The solution method you select in the Define Multi-Point Constraints
dialog box becomes the method used for all features that include MPCs.
These features include, but are not limited to, cyclic symmetry, frictionless
constraints, smart bonding, and user-defined MPCs. For example, if you
want to use the Penalty Method to solve all your analyses involving smart
bonding, you can override the default condensation method by selecting
Penalty Method in the Define Multi-Point Constraints dialog box.
7. Click the OK button to close the dialog box.
8. Run the analysis.

Compatibility Note: Versions 20 through 20.4 SP1


The input used to store the MPC data in versions 20 through 20.4 SP1 is no longer
compatible. To recover the original input, edit the file DS.CST.BAK located in the design
scenario folder (for example, modelname.ds_data\1). This gives the original MPC equations.
The node numbers need to be converted to the corresponding vertex numbers, and then the
equations need to be re-entered through Add Multi-Point Constraint. The format of
the .CST.BAK file is as follows:
#_equations, max_n
#_terms(1)

node(1), DOF(1), M(1)


node(2), DOF(2), M(2)
node(3), DOF(3), M(3)
node(#_terms), DOF(#_terms), M(#_terms)

Constant(1)

Example
Analyze the gear train in Figure 2. Instead of trying to model the gears with beam elements to
mimic the rotational connection, use Multi-Point Constraints. For the gears,
Radius1*rotation1 = -Radius2*rotation2. So the MPC equation is Radius1*rotation1 +
Radius2*rotation2 = 0. For the dimensions and vertex numbers given, the input for the MPC
is as follows:
Equation 1
Constant = 0

Mu
3
9

Figure 2: Gear Train Analyzed with Beam Elements


Pin Constraint

Apply surface pin constraints to cylindrical surfaces to prevent the surfaces from moving or
deforming in combinations of radial, axial, or tangential directions.

1. Select the appropriate surface


2. Right-click in the graphics display and click Add.
3. Select Surface Pin Constraint.
4. In the Creating Pin Constraint Object dialog box, select the appropriate settings:
o Fix Radial: Restricts surfaces so they cannot move, rotate, or deform radially.
Choose this setting to prevent penetration and separation.
o Fix Tangential: Restricts surfaces so they cannot rotate. Choose this setting to
prevent the surfaces from moving tangentially around the circumference of the
pin.

o Fix Axial: Restricts surfaces so they cannot move axially. Choose this setting
to prevent the surface from moving along the length of the pin.
You can also add a des
End Releases

An end release can be applied to a beam element.


An end release allows either or both ends of a beam element to rotate about or translate along
one or more of the local axes of the beam.

Apply End Releases


If you select one or more beam elements using the Selection Select Lines command and
right-click in the display area, you can select the Add pull-out menu and select the Beam
End Releases command to add an end release to each beam element. You can also access this
command via the ribbon (Setup Beam Loads Beam End Release).
Activate the check boxes in the I Node and J Node sections to release the rotation (R) or
translation (T) of that node in one of the local directions. For example, T1 refers to the
translation along the local 1 axis, and R1 refers to the rotation about the local 1 axis. You can
use the buttons to the right to set the releases to certain values.
Tip: The orientation of the beam elements can be displayed using the View
Visibility
Object Visibility
Element Axis commands. Axis 1 points in the
direction from the I Node to the J Node. If axis 1 needs to be reversed for some
elements, this can be done by selecting the elements ( Selection
Select
Lines), right-clicking, and choosing Invert I and J Nodes.

In a normal beam member, the displacement and rotation result is continuous from one
element to the next. Since end releases free a translation or rotation at a node, there is a
theoretical step-change in that result at the node. For example, a rotational end release creates
a hinge. The rotation angle at the end of one beam does not equal the rotation angle at the
start of the next beam. (See figure (a) and (b) below.) However, the FEA results give just one
value of displacement and rotation at a node. Therefore, the effect of the end release is not
visible at the node. (See figure (c) below.) If necessary, add an extra node close to the end
release to see the effect more clearly.

(a) Fixed-fixed beam with a hinge point at 1 and 2.

(b) The theoretical rotation or slope of the beams. Note how the result is discontinuous at the
hinge points.

(c) Since there is only one rotation result in the analysis at the hinge points 1 and 2, the
rotation results are presented as continuous. The length of the step (shown by dashed lines in
the figure) is determined by the size of the element with the end release.
Since the internal shear forces and moments are calculated at each node on each element,
discontinuities can be displayed in these results.
Offsets

Apply an offset to beam elements when the lines comprising the beams cannot be drawn
where the neutral axis is actually located. Offset the beam the appropriate distance and in the
appropriate direction to have the beam properties represented at the correct neutral axis
location. The image below shows how the processor sees the beam elements with beam
offsets applied.

For example, to attach a beam to a plate so that they are properly connected and the forces are
transferred between them, the nodes of the beam elements must be coincident with those of
the plates. Without an offset, this means that the neutral axis of the beam will be at the
midplane of the plate. In reality the neutral axis of the beam is not located there. This is when
an offset must be applied to properly model the assembly. For this example, the neutral axis
of the beam should be located at an offset distance equal to half the height of the beam plus
half the thickness of the plate.

Apply Offsets
If you select one or more beam elements using the Selection Select Lines command and
right-click in the display area, you can select the Add pull-out menu and choose the Beam
Offsets command to add an offset to each beam element. This command can also be accesed
via the ribbon (Setup Beam Loads Beam Offset).
Specify the component of the offset in each global direction in the I Node and J Node
sections. The buttons below can be used to set the offsets to certain values.
Tip: The orientation of the beam elements can be displayed using the View
Visibility
Object Visibility
Element Axis commands. If axis 1 needs to be
reversed for some elements, this can be done by selecting the elements (
Selection
Select
Lines), right-clicking, and choosing Beam Orientations
Invert I and J Nodes.
Important: For nonlinear structural analyses, the user must activate the Large
rigid body rotation option in the Advanced tab of the Element Definition
dialog box to get the most accurate results from the beam offsets.

Contact Pairs

Keep in mind that loads are transferred from one element to the adjacent element only if the
nodes are connected. Typically, the elements within one part are connected directly through
the nodes since the mesh is continuous.
How different parts in the model are connected is determined by the Contact heading in the
browser (tree view). When two surfaces of different parts touch, they will be connected
together based on the default contact type, unless a specific contact pair is created and
defined as a different type of contact. See the Types of Contact page for an explanation of the
available contact types and the mesh requirements.

Generate Contact Pairs


Specific contact pairs need to be created and assigned to the appropriate contact type when
the default contact type is not appropriate for the pair.
Automatic Contact Pairs
When working with a CAD model, you can choose whether the interface automatically
creates the list of all contact pairs in the model, or whether you want to create the contact
pairs manually. This choice is set from the
Options CAD Import tab; Global
CAD Import Options button, then set the Automatically generate contact pairs option
accordingly.

Automatically generate contact pairs activated. When a CAD model is


opened in the interface, the surfaces that are physically in contact (zero
gap) are listed in the Contact branch of the browser. The contact type for
each pair is Default so that they follow the setting of Contact (Default:). To
set certain contact parameters to be used between specific surfaces,
select the contact pair in the browser, right-click, and choose the new type
of contact.

Automatically generate contact pairs not activated. When a CAD model is


opened in the interface, the time required to detect all of the surfaces that
are physically in contact is not expended. You can manually choose parts
or surfaces to be in contact as described next.

Add Additional Contact Pairs


Regardless of whether the contact pairs are generated automatically or not, you can create
additional contact pairs. This may be needed because you want contact between two parts
(instead of between the individual surfaces on the parts), or the two surfaces are not initially
in contact, or because the original entry was deleted. Also, the contact pairs in a hand-built
model must be specified manually.
There are two methods to generate the additional contact pairs:

Create pairs one at a time. This method is appropriate for CAD and handbuilt models. First select two surfaces, two parts, or one surface and one

part in the browser. Once they are selected, right-click the heading for one
of the surfaces or parts and select the Contact command. Select the
appropriate type of contact for this pair. A new heading will be created
under the Contact heading in the browser for this pair. Alternatively, the
parts and surfaces can be selected in the display area, then right-click and
choose Contact. Normally, which surface is selected first and which is
selected second does not matter, other than controlling the order of the
listing in the browser. For surface-to-surface contact in a nonlinear
analysis, the order may be important when defining the Contact type as
Point to Surface. The first part or surface selected becomes the primary
surface, and the second selection becomes the secondary surface. (See
the Surface-to-Surface Contact Options page for details on the Contact
Type.)

Create multiple contact pairs at a time. This method is appropriate only for
CAD models. First select two or more parts in the browser or in the display
area. Then, right-click and select Contact Create Contacts Between
Parts. Any pair of surfaces that are in contact between the selected parts
will be listed in the Contact branch of the browser. The type of contact for
each pair can then be changed individually by right-clicking on the entry in
the browser.

Contact between different surfaces of the same part is not permitted except in nonlinear stress
surface-to-surface contact. (Use the Contact tree to generate a contact entry within a single
part, but the processors will not handle it.)
Sort Contact Pairs
When contact pairs are added to the model, they are listed at the bottom of the contact branch
of the browser in the order that they are created. The contact pairs will be sorted whenever
the model is opened again. There are three different types of contact entries:

Part A Surface B with Part C Surface D (A/B with C/D in browser)

Part A with Part C Surface D (A with C/D)

Part A with Part C (A with C)

The sorting order of the contact list is by A, then C, then B, and then D
(ascending).

Color of Contact Pairs


Contact pairs that are created automatically (or prior to version 23) use black text in the
browser. Contact pairs that you create explicitly use blue text in the browser.
Tip: Selecting a contact pair in the browser highlights the corresponding lines in
the model. To easily see a contact pair, right-click on the heading for the pair that
is of interest and choose the Isolate command from the context menu. The pair
will be highlighted and all other surfaces in the model will be hidden.
Alternatively, to view the model without shading, use View
Appearance
Visual Style
Edges or View
Appearance
Visual Style
Mesh. If

model colors make it difficult to see highlighting, use View


Appearance
Color By
Part or View
Appearance
Color By
Surface to change the
display. In addition, the highlight color can be set from the Setup tab under the
Options dialog.

Consider surface contact between three different parts where the materials of these parts will
create different static coefficients of friction. You must create two contact pairs and assign a
static coefficient of friction for each pair. You would not be able to set the default contact type
to surface and set up two different friction coefficients.
As another example, consider a model has 50 surfaces in two different parts that interact with
each other. Assume that various contact pairs have different types of contact. Then, selecting
entries and changing the contact type for each pair may be the best approach. However, if
only two pairs of surfaces are in contact and the other 48 pairs of surfaces are considered to
be bonded, then you should set the default for the entire model to Bonded. The other two
surfaces would then be selected and set to Surface contact. The idea is to minimize the
number of contact pairs that must be defined. Specify the default contact type that is most
prevalent within the model.
Order of Contact Affects the Model
If the exact same contact pair is created (that is, duplicate contact pair), only the first entry in
the browser is used. Subsequent entries are ignored. When both part to part and part/surface
are defined at the same face, the part/surface entry controls the contact type for that pair, and
the other surfaces that are in contact follow the part setting. See Figure 1 for examples.

(a) The same part/surfaces are defined in contact. Free contact is the result because it is listed

(c) The contact between part 1 surface 6 and part 2 surface 5 (1/6 with 2/5) is a smaller more d
Topics in this section

Types of Contact
The FEA Editor always keeps separate vertices (duplicate vertices) at the boundaries between
different parts. That is, two or more vertices exist at the same coordinate: one for each part.
The type of contact determines whether:

the separate vertices create one node shared by the parts (bonded
contact),

two separate nodes are created but not connected together (free contact),
or

two separate nodes are created and connected together with an


automatically-generated element (surface contact).

The merging of the vertices occurs when a Check Model operation is performed or at the
beginning of the solution phase (Run Simulation). Depending on the analysis type, element
type, and contact type, the mesh may need to be matched between the contacting parts.

Mesh Requirements for Contact


The following table indicates whether adjacent parts within an assembly need to be modeled
with a zero-gap fit, or if an initial gap or interference is allowed between parts for the various
analysis and contact types:
Analysis Category
Linear
Nonlinear

Analysis Category

Thermal
Electrostatic

The following table indicates whether or not the meshes need to be matched between adjacent
parts for different analysis and contact types:

Analysis Type
Linear, Static Stress
Linear, Natural Frequency (Modal) and Natural Frequency (Modal) with Load Stiffening
Linear, Critical Buckling Load and Transient Stress (Direct Integration)
Linear, Other
Nonlinear
Thermal
Electrostatic

Note 1: If using smart bonding, the mesh can be matched or unmatched. If not using smart
bonding, then the mesh must be matched to be bonded. Smart bonding is set under the
following dialog boxes. See the referenced pages for details on smart bonding, including the
limitations on the element types.

Linear static stress: Setup Model Setup Parameters Contact


tab Enable smart bonded/welded contact. See the page Analysis
Parameters: Static Stress with Linear Material Models.

Linear modal: Setup Model Setup Parameters Contact tab


Enable smart bonded/welded contact. See the page Analysis
Parameters: Natural Frequency -Modal.

Linear modal with load stiffening: Setup Model Setup Parameters


Contact tab Enable smart bonded/welded contact. See the page
Analysis Parameters: Natural Frequency -Modal with Load Stiffening.

Linear critical buckling: Setup Model Setup Parameters Contact


tab Enable smart bonded/welded contact. See the page Analysis
Parameters: Critical Buckling Load.

Linear transient stress: Setup Model Setup Parameters Contact


tab Enable smart bonded/welded contact. See the page Analysis
Parameters: Transient Stress -Direct Integration.

Thermal/heat transfer: Setup Model Setup Parameters Contact


tab Enable smart bonded/welded contact. See the page Analysis
Parameters: Steady-State Heat Transfer or Transient Heat Transfer.

Electrostatic: Setup Model Setup Parameters Contact tab


Enable smart bonded/welded contact. See the page Analysis
Parameters: Electrostatic Current and Voltage or Electrostatic Field
Strength and Voltage.

Note 2: The following analysis types use the results from the modal analysis. Therefore,
these analyses can take advantage of smart bonding (see Note 1) without any additional
contact setup required:

Response Spectrum

Random Vibration

Frequency Response

Transient Stress (Modal Superposition)

DDAM.

Note 3: The mesh can be unmatched if using tied surface contact to perform the bonding, in
which case the contact type would be Surface. If not using tied surface contact, then the mesh
must be matched, and the contact type is Bonded. Tied surface contact is set under the
Settings for the contact pair. See the Surface-to-Surface Contact Options PAGE for details on
tied contact.
Note 4: For linear analyses, the Surface Contact column refers to three available contact
typesSurface, Slide/No Separation, and Separation/No Sliding, all of which are selectable
from the browser. Only the first type (Surface) is applicable to thermal analyses. Thermal
Surface contact can optionally include thermal resistance, which results in a temperature
differential between the contacting surfaces when heat is flowing. Surface contact types and
options are set up differently for nonlinear analyses. Refer to the Surface-to-Surface Contact
page and the Options sub-page for more information.
Note 5: The Shrink Fit Contact column refers to two available contact typesShrink
Fit/Sliding and Shrink Fit/No Sliding. The amount of interference between parts can be
represented in the CAD geometry (by overlapping the parts). Upon importation of the model
into Simulation Mechanical, the surfaces are matched (zero interference, zero gap). That is,

the interference is removed from the geometry and the mesh is matched between the parts.
However, the amount of interference is retained in numerical form. For further discussion of
this, see the Shrink Fit / Sliding and Shrink Fit / No Sliding sections below.
Important: Note that a matched mesh requires the two surfaces in a Simulation
Mechanical model to be touching (zero gap between them). However, a small
interference or gap may be present in the original CAD assembly, and this is
desirable in some situations.

A small gap in a CAD model can be removed during meshing by using the
mesh matching tolerance. Removal of the gap is necessary to match the
nodes. See the paragraph, Mesh matching on the, Model Mesh Settings:
Model page.

If a gap between parts is critical to the analysis (that is, you do not want to
remove the gap), then use manually created gap elements (linear static
stress) or surface to surface contact (Mechanical Event Simulation).

Interferences between parts of a CAD assembly will be removed during


CAD data importation. However, the interference is recorded numerically,
as discussed in Note 5 above.

The type of contact specified by the top-level Contact (Default: __) setting in the tree view
is applied to any two surfaces and are either not listed individually, or are listed individually
but set as Default contact type. For example, if it can be assumed that all parts of an
assembly are in perfect connection with each other able to transmit all types of loads at
their interface then you can leave the default contact set to Bonded.
Each of the available types of contact are covered in the following sections.

Default
If the Default command is selected for a contact entry, the type of contact will follow the
global Contact (Default: __) setting. Default can only be selected in the tree view for an
existing contact pair; it cannot be selected when creating the contact pair.

Bonded
Bonded contact is applicable to all element types. The two surfaces will be in perfect contact
throughout the analysis when bonded, and the loads are transmitted from one part to the
adjacent part. In a stress analysis, when a node on one surface deflects, the node on the
adjoining surface will deflect the same amount in the same direction. In a heat transfer
analysis, the same temperature occurs in each part at the connection, and so forth.
Smart bonding is a method to connect the nodes on adjacent parts even though the meshes do
not match. (See the Notes for Table 1 to activate the smart bonding and limitations on
element types and analysis types.) Since the nodes may not line-up, a tolerance is required to
match the mesh from part A surface B with the mesh from part C surface D. When using
smart bonding, right-click a contact pair in the tree view and choose Settings to access the
tolerances. The two options for the Tolerance type are as follows:

Fraction of mesh size uses the input in the Value field times the mesh
size to create a dimension. Nodes on the one surface within this distance
of the other surface will be bonded. (The mesh size is based on the
average element size at the location of the node.)

Absolute size uses the dimension in the Value field for matching the
nodes to the adjoining surface.

When the contact type is defined as part A with part C or part A with part C surface D, any
nodes on part A within the tolerance of part C surface D will be bonded with the smart
bonding. This may affect what size tolerance you want to use. (Technically, the software
determines which surface is bonded to the adjacent surface. It does not depend on the order of
the entry shown in the tree view.)
Note: Do not use the smart bonding option if the same mesh will be used for
another analysis and if the other analysis does not support smart bonding (such
as a nonlinear analysis when using Welded contact). Or, if the contact type will
be changed to a type that does not support smart bonding (such as from Bonded
to Surface contact), then the meshes should be matched.

Smart Bonding Function


Smart bonding transmits loads from the nodes on one part to the adjacent nodes on the other
part by using equations to link the degrees of freedom (displacements for stress analysis,
temperature for heat transfer) of the parts together. This is done with the same method as
multi-point constraints. (For details on user-defined multi-point constraints, see the MultiPoint Constraints page.) In stress analysis, the translations are linked for brick and 2D
elements. For plate to plate elements, the translations and rotations are linked. See Figure 1
for examples.
Attention: When using smart bonding to bond plate elements to brick elements
in a stress analysis, the rotation of the plate nodes is set to zero. (See Figure
1(b).) This is equivalent to applying an RxRyRz boundary condition to the nodes
of the plate.

Since the nodes on surface D may be connected to nodes on surface B outside the area of
surface D, the smart bonding can distribute the load from part C to part A over a larger area
than the physical contact area. Naturally, the degree that the load is spread out depends on the
size of the mesh. See Figure 1(a).

TX26 = TX22 + (a/b)(TX23-TX22)


TY26 = TY22 + (a/b)(TY23-TY22)
TZ26 = TZ22 + (a/b)(TZ23-TZ22)
T is translation in X, Y, or Z at
indicated node.

(a) Solid (P1) to solid (P2) connection with smart bonding, stress analysis.
Since part 2 is connected to part 1 at nodes 22 and 23, the load is transferred over a larger
area than the interface between parts 1 and 2. This approximation is more accurate as the
mesh size on part 1 is reduced.
TY26 = TY22 + (a/b)(TY23-TY22)
TZ26 = TZ22 + (a/b)(TZ23-TZ22)
RX26= 0
R is rotation about X, Y, or Z at
indicated node.

(b) Plate (P2) to solid (P1) connection with smart bonding, stress analysis.
(2D view considered for simplicity)

T26 = T22 + (a/b)(T23-T22)


T is the temperature at the indicated
node.

(c) Solid to solid connection with smart bonding, heat transfer analysis.
Figure 1: Smart Bonding Examples
Smart Bonding Tip

Use smart bonding to connect the midside nodes on part A (a part where more accurate results
are desired) with the corner nodes on part B (a part without midside nodes, not requiring as
high of accuracy).

Without smart bonding, the midside


nodes in the part on the right can pull
away from the face of the left part.
(Only nodes on contact face shown for
clarity.)

With smart bonding, the midside nodes


are bonded to the face of the left part.
Thus, the midside nodes cannot pull
away. (Parts shown with slight gap for
clarity.)

Figure 2: Bonding Midside Nodes


Smart Bonding Note

Models created prior to the introduction of smart bonding may have relied on an un-matched
mesh to create free surfaces even though the contact type was set to bonded. If such models
are re-analyzed, the smart bonding should not be enabled; otherwise, such surfaces will now
behave as bonded. See Figure 3. (By default, smart bonding is disabled for legacy models.)

(a) Prior to smart bonding, parts are not bonded because the nodes do not match. (Actually, no type of contac

(b) If the same model is opened in software with smart bonding capability, and if the contact type is bonded, a
Figure 3: Free Contact with Unmatched Mesh (Older Models)

Spot welding is an example of congruent surfaces where some nodes do not match and you
do not want these nodes to be bonded.
Smart Bonding:
The above discussion (Figure 1) implied that the nodes on part 2 are connected to the nodes
on part 1. You do not have this level of control. Instead, you specify whether the nodes on the
surface with the coarser mesh are connected to the other side, or vice versa. This selection is
made from the Setup Model Setup Parameters dialog box. Beyond the user-selection,
the software decides which surface is the finer mesh and which is coarser. Note that all
elements on the contacting surface are considered; the mesh size at a portion of the contact
surface is not the deciding factor between fine and coarse. Only nodes that are within the
tolerance (see above) of the opposite surface are used; nodes outside the contact area are not
used to determine fine or coarse. For example, if you define a contact pair between parts 1
and 4, then all nodes that contact between parts 1 and 4 are counted regardless of whether the
contact is at one face or many faces; the part with the fewer contact nodes becomes the
coarser mesh. (The same is true for the default contact setting; all contact nodes between two
parts are used to determine whether the mesh is fine or coarse.) For contact pairs that define
contact between part/surface and part/surface, the nodes that contact on the specified surfaces
are used to determine which mesh is fine or coarse.

Tip: Using a matched mesh always gives more accurate answers. This is true
whether node bonding or smart bonding is used to bond the surfaces together.

When choosing what type of smart bonding to use (Coarse bonded to fine mesh or Fine
bonded to coarse mesh), it is helpful to consider what is supposed to occur at the interface
between the two parts. This is demonstrated in figures 4 through 7.
For stress analysis where the interface experiences pure compression, all of the nodes on the
interface should deflect the same amount. Thus, choosing Fine bonded to coarse mesh will
give more accurate results. See Figure 4.

(a) The Model. A simple compression model.

(b) Incorrect. When using Coarse bonded to fine mesh smart bonding, only the nodes on the coarse mesh ar

(c) Correct. When using Fine bonded to coarse mesh smart bonding, the nodes on the fine mesh are forced to
Figure 4: Stress Analysis Under Pure Compression

For stress analysis where the interface experiences bending, the nodes on finer mesh surface
must be allowed to move independently of the nodes on the coarser mesh; that is, both sets of
nodes should displace so as to create the same radius of curvature. Thus, choosing Coarse
bonded to fine mesh will give more accurate results. See Figure 5.

(a) The Model. A simple bending model.

(b) Incorrect. When using Fine bonded to coarse mesh smart bonding, the nodes on the fine mesh are forced

(c) Correct. When using Coarse bonded to fine mesh smart bonding, only the nodes on the coarse mesh are
Figure 5: Stress Analysis Under Pure Bending

In heat transfer analysis, the flow of heat should be continuous from the finer mesh into the
coarser mesh. Thus, choosing Fine bonded to coarse mesh will give more accurate results.
See Figure 6. Similar philosophy applies to electrostatic analysis.

(a) The Model. A simple two-part model with the bottom surface exposed to a hot environment and the top su

(b) Incorrect. When using Coarse bonded to fine mesh smart bonding, only the nodes on the coarse mesh ar

(c) Correct. When using Fine bonded to coarse mesh smart bonding, the nodes on the fine mesh are forced t
Figure 6: Heat Transfer Analysis
Another consideration in choosing the type of smart bonding is that there needs to be nodes
on the chosen surface that physically match the opposite surface. See Figure 7.

Coarse bonded to fine mesh smart bonding versus Fine bonded to coarse mesh smart bonding. If only nod
Figure 7: Bonding Drastically Different Mesh Sizes

Keep in mind that smart bonding approximates the transfer of loads at the boundaries of two
parts. In a stress analysis, the continuity of the displacements is conserved. Other quantities
such as force, stress, derivative of the displacements are not continuous. Therefore, the results
at the boundary where smart bonding occurs may be inaccurate depending on the relative
mesh density. However, results remote from the boundary should be accurate, relative to the
effects (if any) caused by the smart bonding. If accurate results are desired at the boundary
between the parts, smart bonding should not be used; the meshes should be matched instead.

One advantage of smart bonding is to perform what-if studies and change the mesh on one
part without the need to re-mesh the entire model. The smart bonding will maintain the
connection at the contact faces even though the mesh may not align. See Figure 8.

Figure 8: Example Use of Smart Bonding.

A mesh sensitivity study can be performed by changing the mesh on part 2 [P2] without the need to re-mesh t
The equations created with smart bonding are solved using a penalty method. A stiffness is
applied to the MPC equations, and this stiffness is added to the normal stiffness matrix. The
stiffness of the penalty equations depends on the Solution method you choose (see MultiPoint Constraints). For the Penalty Method you enter the stiffness in the Penalty multiplier
field. A multiplier of 0 implies that the parts are not connected, and a multiplier of infinity
implies a perfect bonding between the parts. Unfortunately, infinite stiffness is not acceptable
in a numerical solution, nor is it required. Generally, a penalty multiplier on the order of 1E4
to 1E6 provides a satisfactory solution. However, some analyses may give a maximum to
minimum stiffness warning (max/min stiffness) or fail to find a solution if the penalty
multiplier is too large.
To judge the accuracy of the solution, the summary file includes a line with the satisfaction
factor. A value of 100% indicates that the MPC equations (such as in Figure 1) are satisfied
accurately. Any value less than 100% indicates some separation of the parts. It is not possible
to say that X% indicates the solution is wrong. Rely on experience to judge when the
satisfaction indicates an unacceptable result.

Welded
Welded contact is applicable only to brick and 3D element types. If the Welded command is
selected, the nodes along the edges of the contact surfaces act the same as if the Bonded
command were selected. The nodes along the interior of these surfaces act the same as if the
Free/No Contact command were selected.

Free/No Contact
Free/No Contact is applicable to all element types. If the Free/No Contact command is
selected, the nodes on the two surfaces in this contact pair will not be collapsed to one node
even if the mesh is matched. In a CAD model, the mesher will not necessarily force the mesh

to match. These nodes will not transmit a load between the parts. In a stress analysis, the
nodes will be free to move relative to the nodes on the other surface. In a heat transfer
analysis, no heat will be conducted between the parts, and so forth.

Surface Contact
Surface contact is applicable only to the element types indicated below. Surface contact is
available for CAD solid models, 2D created meshes, and hand-built models as follows.

Linear static stress: Surface contact between brick parts, between brick
and plate parts, and between 2D parts can be specified. Contact is
created only if the gap between the parts is zero. In the case of
brick to plate, set the contact direction based on the brick part. See the
next section for setting the contact direction. Automatic contact between
plate parts is not supported. (Naturally, you can create gap elements
manually between any type of elements.)

If the Surface Contact command is selected, a zero-length contact element (similar to


a user-created gap element) is placed between the nodes. The nodes will be free to
move away from each other, but the nodes cannot pass through each other when they
come into contact. Imagine a very small line created between the nodes on these
surfaces. If that line becomes longer during the analysis, it will have no effect on the
model. If that line becomes zero length, it will act as a spring with a stiffness value
that will resist this motion.
When using Surface Contact, the analysis will involve an iterative process; hence,
the analysis will take longer to run than an analysis with bonded contact. This process
will be used to determine if the deflection due to the loading will cause each pair of
nodes on these surfaces to be in contact or not.
Note: For additional modeling information related to contact elements in a
linear stress analysis see Performing Analyses with Gap Elements.

When the contact elements are defined for surface contact pairs, a direction is
calculated for each contact element. The method used for calculating this direction
can be specified by right-clicking on the heading for the surface contact par in the tree
view and selecting the Settings command to access the Contact Options dialog. The
direction calculation method can be selected in the Surface contact direction dropdown box. The default option of Calculate by matching directions will calculate the
direction from the normals of the elements at each node. This is the recommended
method. If a mesh is coarse, some of the elements may be non-flat (i.e. the fourth
node is not coplanar with the other three). In this case the default method may not
result in valid directions. To achieve valid directions, you can increase the value in the
Direction tolerance angle field. If one of the surfaces does contain flat elements, you
can correct the direction calculation by selecting either the Normal to the first
part/surface or Normal to the second part/surface option in the Surface contact
direction drop-down box. You should select the option for the part/surface that
contains flat elements.

Friction can be added to a contact pair or the default contact if Surface Contact or
Edge Contact is selected. To turn on friction, right-click the contact pair and then
select the Settings command. This will open a dialog that will allow you to activate
the Include friction check box and define the Static friction coefficient. Contact in
Static Stress with Linear Material Models analyses assumes that the contact force is
transmitted through the same set of nodes as originally in contact and the direction of
the force is the same as the original normal direction. That is, the deflection does not
change the points in contact or direction (small deformation theory).
In Figure 9, a block is on a fixed surface. A normal force, N, is applied to allow the
generation of a friction force, F f .A static coefficient of friction of exists between the
block and the fixed surface. A lateral force, F is applied to the block. A spring is
placed between the block and a wall to create a statically stable model. F f can range
from 0 to N as follows:
For F <= N:
Ff=F
Fs=0
For F >= N:
Ff=0
Fs=F

Figure 9: Static Friction


In summary, once the force exceeds the maximum value of the friction force, the
friction force goes to zero and no longer resists the motion. If what happens after this
is important, a Mechanical Event Simulation (MES) analysis using surface to surface
contact should be performed. In MES, a dynamic coefficient of friction can be defined
to calculate the friction force after the static friction force is exceeded.

Mechanical Event Simulation and nonlinear stress: Contact between


combinations of brick, shell, 2D, beam, and truss parts can be specified
with the surface-to-surface contact. Because this type of analysis
calculates large deflections, the parts do not need to be physically in
contact to create automatic contact.

The contact parameters in a Mechanical Event Simulation/nonlinear static stress


analysis are entered in the surface to surface contact screens. Both static and sliding
(dynamic) friction can be included in the analysis. For details, see the Surface-toSurface Contact page.

Heat transfer analysis: Surface contact between brick parts, between brick
and plate parts, between plate parts, and between 2D parts can be
specified. Automatic contact is created only if the gap between the
parts is zero.

Surface contact in a thermal analysis creates a zero-thickness 2D or 3D element


between the surfaces. Since the nodes on the two parts are not merged together, there
will be a temperature gradient across the joint.
For a thermal analysis, a contact resistance can be defined for any surface contact pair.
This can be used to simulate imperfect contact resulting in a temperature gradient
between the parts. This can also be used to simulate a thin film between two parts.
You can define the resistance by right-clicking on the heading for the contact pair and
selecting the Settings command. You can choose Total Resistance or Distributed
Resistance in the Thermal Resistance: drop-down box. If the Total Resistance
option is selected, the resistance value defined in the Value: field will be evenly
distributed over the surface pair. (For 2D planar models, the total resistance is for the
thickness of the mating parts. For 2D axisymmetric models, the total resistance is for
1 radian.) If the Distributed Resistance option is selected, the value defined in the
Value: field will be divided by the area of the contact pair and applied to the model.
Thermal Contact Example
Contact elements are not created at boundaries where two parts with surface contact
are bonded to a different part (or multiple parts). For example, imagine parts 2 and 4
are defined as surface contact with each other and bonded to part 1. See Figure 10(a).
A detail of the joint shows that parts 2 and 4 are separated from each, and the gap is
filled with an 8-node thermal contact element (for 3D models) or 4-node thermal
contact element (for 2D models). See Figure 10(b). However, no contact element is
created at the location where parts 2 and 4 are bonded to part 1.

(a) Example three part model. Parts 2 (P2) and 4 (P4) have surface contact defined between them. Bot

contact element

(b) Detailed view of the contact between the parts. Thermal contact elements are create everywhere ex
Figure 10: Example Thermal Contact
Tip: A thermal resistance of 0 --- indicating a perfect contact between the
parts --- is not a good choice. A resistance of 0 implies a conductance of
infinity, which leads to inaccuracies during the numerical solution. Use a
small, nonzero value to simulate such conditions when they exist.

Sliding / No Separation
Bonds contact surfaces in the normal direction while allowing the surfaces to slide relative to
each other in the tangential direction.

Separation / No Sliding
Contact surfaces can partially or fully separate in the normal direction, but cannot slide
relative to each other in the tangential direction.

Edge Contact
If the Edge Contact command is selected, the nodes along the edges of the contact surfaces
will act the same as if the Surface Contact command were selected. The nodes along the
interior of these surfaces will act as if the Free/No Contact command were selected.

Shrink Fit / Sliding


Behaves like Surface Contact but with initial interference between the contacting parts. The
Contact Options dialog box includes an Interference input field and an Automatic checkbox
to the right of this field. If your model is CAD-based and there is geometric interference in
the CAD assembly, the interference is removed from the geometry and the meshes are
matched to each other in a net fit (zero-interference, zero-clearance). However, the
interference from the original CAD assembly is retained numerically. You will still have to
define shrink fit contact in Simulation Mechanical (assuming you want to quantify the effects
of the interference). The Automatic interference option is activated by default, and the
interference magnitude is assigned automatically on a node-by-node basis (that is, it can be
non-uniform).
Important: In order for Automatic interference assignment to work correctly, you
must define contact by surface pairs. Automatic interference will not work for
part-to-part contact or for the Default (global) contact type.

If you disable the Automatic option, or if you are using a hand-built model or a CAD model
without geometric interference, you can specify a uniform radial interference between the
contacting parts. For example, if a bearing with a bore of 40 mm diameter is pressed onto a
40.03 mm diameter shaft, the radial interference is 0.015 mm [half of the diametral
interference, or (40.03-40) / 2]. Manually specifying the interference is a quick way to
determine the effects of various intensities of fit without having to modify the CAD
geometry. It also makes it easy to determine the effects of interferences when they are not
represented in the CAD model or for hand-built models.
A CAD assembly can be modeled with a net fit between the contacting parts. A hand-built
assembly must be modeled with matched surfaces (net fit). In either case, a uniform
interference can be manually specified within the Contact Options dialog box. In addition, the
geometric interference, if any, in a CAD model can be overridden with a user-specified,
uniform interference value.

Shrink Fit / No Sliding


Behaves like Separation / No Sliding contact but with initial interference between the
contacting parts. Use this option when the intensity of fit and friction are great enough to
prevent relative motion (sliding) between the contacting parts. The radial interference is
specified automatically (CAD-based) or manually in the same manner as it is for Shrink Fit /
Sliding contact (see above).

Example
The following images illustrate two contact types. Note the box enclosing the end
of the gun in the first image. In this region, the outer diameter of the metal ring

has an interference fit with the caulk gun frame. The second image shows a
cross-section of the frame-ring interface.

The contact type for interface 1 is Surface Contact. The contact does not
constrain the ring from moving in either the normal or tangential
directions.

The contact type for interface 2 is Shrink Fit / No Sliding. The contact has
an initial interference fit as the outer diameter of the ring is larger than
the inner diameter of the frame. However, the ring is still free to move in
the normal and tangential directions if the friction force from the shrink fit
is overcome.

Troubleshooting: Contact Examples

Different types of issues can occur with the connection between adjacent parts. One issue is
that the nodes might not be connected together as intended or desired. When parts comprised
of two different element types are connected, the element type with the fewest degrees of
freedom (DOF) at each node dictates the behavior of the connection. For example, elements
with no rotational DOF (such as bricks and tetrahedra) cannot resist moments. So, these
connections have no rotational stiffness. The supported nodal degrees of freedom dictate what
type of load is transmitted between the parts. This page discusses a couple of these issues and
provides methods for producing the desired connection characteristics.
Note: Unless a description on this page specifically refers to smart bonding, the
description is for contact that requires the meshes to be matched.

Bonded Contact Between Plates and Bricks:


The issue here is related to the fact that the degrees of freedom for each part are different.
Bricks have only the three translational degrees of freedom (Tx, Ty, and Tz) while plates (and

shells) have all DOF except for rotation about the direction normal to the element. Therefore,
the bricks cannot apply a moment to the plates to prevent the plates from rotating. Without
special considerations, a connection between plates and bricks along a straight line of nodes
will result in a hinge. This is depicted in Figure 1. In the case of linear static stress, hinge
connections frequently result in a model that is not statically stable, though the example
pictured below is stable due to the constraint at the opposite end of the plate.

Figure 1 Hinged Combination of Bricks and Plates


The black arrows represent the hinge line. The schematic on the right shows a side view and
how the plate deflects; note the rotation at the hinge line. If the free end of the plate was not
supported, the plates is free to rotate around the hinge line which would result in a statically
unstable configuration.
To avoid this hinge effect, the bonded contact between the plate and brick needs to be
distributed over an area; specifically, at least three nodes not in a straight line. See Figure 2.

Figure 2 Statically Stable Combination of Bricks and Plates


In the top method, the plate is embedded one element deep into the bricks. (The elements are
shrunk to show the space between them.) This method is applicable to hand-built (structured
mesh) models. In the bottom method, a tee connection of plates is created on the surface of
the bricks. In the schematics on the right, which show the side view, the arrows represent the
reactions at the brick/plate interface. (These are in fact the forces that are transmitted between
the nodes of the bricks and plates.) Since force couples can resist moment loads, the plate is
now statically stable.

Bonded Contact Between Bricks and Beams:


The connection between beams and bricks has a problem similar to the plate-to-brick
connection: the bricks cannot prevent the beams from rotating. Beams support all six degrees
of freedom (three translations and three rotations), but the bricks only support the three
translational DOF (Tx, Ty, and Tz). The solution is likewise similar to plate/beam
connections: make a web of beams (or spokes) to connect to the bricks. The beams must
connect to the bricks at three or more points, not in a straight line, to form a statically stable
solution. A beam attached to a brick at a single point acts like a ball joint, which typically
results in statically unstable models. Figure 3 shows an example.

Figure 3 Statically Stable Combination of Bricks and Beams


The left side of the figure shows the beams and bricks together; the middle shows the beams
only. Note how the vertical member is connected to four other beams that tie to the bricks at
multiple points. If a torque were applied to the beams, such as in the schematic on the right
side, then the black arrows represent the reactions provided by the bricks.
Note: There are a number of ways to automate the creation of the spoke beams,
depending on the application. For applying a load (such as a torque), select the
surface, right-click, and select Add
Create Remote Load. If working with a
CAD model, use Draw
Design
Centroid Creator to create a construction
vertex at the centroid of the surface. See the Remote Loads and Constraints and
Add Geometry pages, respectively.
Tip: In older versions of the software, attaching beams to brick parts was one
way to apply a moment to the brick part. Another method was to add two or
more nodal forces with tangential directions to the perimeter of the brick surface.
In the current software, you can apply moment loads directly to surfaces of brick
parts. See the Moments page for more information.

Bonded Contact Between Plates and Beams:


There are two situations to consider when combining plates with beams.
1. A plate cannot resist the moment about the axis normal to its face.
Therefore, if you attach a beam normal to a plate, the plate element will
not be able to keep it from rotating. To avoid this, it is recommended to
add beams (spokes) from the end of the normal beam to a few other
nodes in the plate to form a statically stable model. The method is the
same as for beam/brick connections, as shown in Figure 3.
2. If you model a beam lying on a plate, you must remember that you are
drawing the midplane of the plate elements and the neutral axis of the

beam elements. Therefore, if you draw beam part along the plate, you are
really saying that the neutral axis of the beam lies along the midplane of
the plate. This is most likely not accurate. There are two possible ways to
model this situation correctly within the FEA Editor environment.
o

You can draw the neutral axis of the beam and the midplane of the
plate where they actually exist and connect them in several places
with effectively rigid beams.

You can draw the neutral axis of the beam along the midplane of
the plate. Then, using Selection
Select
Lines, select the
beam elements. Right-click in the display area and select Add
Beam Offset from the context menu (or click the Setup
Beam
Loads
Beam Offset ribbon command). Specify the distance and
direction between where the beam is drawn and where the neutral
axis is actually located. For example, if an 8 inch high wide flange
beam is connected along a 0.5 inch thick plate, the offset distance
would be 4.25 inches in the appropriate direction (half the plate
thickness plus half the beam height).

Analysis-Specific Information

Important: This section of the help does NOT provide a comprehensive compilation of
analysis-specific setup topics. Please refer first to the General Information (Common to
Multiple Analysis Types) content. Any topics (loads and constraints, product interoperability
information, modeling considerations, and so on) that are common to two or more different
analysis type groups are contained in the General Information section.
The topics contained in this Analysis-Specific Information section, and listed below, are those
that are only applicable to a specific family of analyses (Linear, Nonlinear, Thermal, or
Electrostatic). In addition, qualifying information may be provided for topics that are covered
in the General Information section but that are treated somewhat differently in a particular
analysis-specific context.
Topics in this section
Linear Analyses

In this section, we cover aspects of setting up and performing the analysis that are specific to
linear analyses. Much of the setup process, types of elements, material models, analysis
parameters, and more are also common to other analysis types, such as nonlinear analyses.
Therefore, first check the General Information (Common to Multiple Analysis Types) section
of the Help. If an item pertains to two or more analysis categories, it is covered in the
General Information section. Otherwise, if an item only pertains to a specific analysis type, it
is covered within the Analysis-Specific Information branch.
In some cases, a particular item may have characteristics that are common across multiple
analysis types, but there are also analysis-specific differences. In such cases:

Common characteristics are covered in the General Information (Common to Multiple


Analysis Types) section.

Analysis-specific characteristics are addressed in one or both of the following ways:


o A qualifying statement is included in the General section, and/or
o Differences are covered in one of the analysis-specific branchesLinear,
Nonlinear, Thermal, or Electrostatic.

Beam Elements

A beam element is a slender structural member that offers resistance to forces and bending
under applied loads. A beam element differs from a truss element in that a beam resists
moments (twisting and bending) at the connections.
These three node elements are formulated in three-dimensional space. The element geometry
specifies the first two nodes (I-node and J-node). The third node (K-node) is used to orient
each beam element in 3D space (see Figure 1). A maximum of three translational degrees-offreedom and three rotational degrees-of-freedom are defined for beam elements (see Figure
2). Three orthogonal forces (one axial and two shears) and three orthogonal moments (one
torsion and two bending) are calculated at each end of each element. Optionally, the

maximum normal stresses produced by combined axial and bending loads are calculated.
Uniform inertia loads in three directions, fixed-end forces, and intermediate loads are the
basic element based loadings.

Figure 1: Beam Elements

Figure 2: Beam Element Degrees-of-Freedom


Note: The mass moment of inertia about the longitudinal axis, I1, is
approximated for beam elements. Specifically, the HRZ lumping method is used
to generate a lumped mass matrix from the consistent mass matrix. The total
translational mass of an element is preserved, while the rotational mass is
approximate.

For rotation about axes 2 and 3, only the mR2 effect is considered, where R is the distance
from the rotation point to the element. The mass moments of inertia, I2 and I3, are calculated
based on the slender rod formula (I2 = I3 = ML2/12).
The three mass moments of inertia only impact Natural Frequency (Modal) and Natural
Frequency (Modal) with Load Stiffening analyses."

Figure 3: Mass Moment of Inertia Axes

Use Beam Elements When

The length of the element is much greater than the width or depth.

The element has constant cross-sectional properties.

The element must be able to transfer moments.

The element must be able to handle a load distributed across its length.

Part, Layer, and Surface Properties for Beam Elements


The following table describes what controls the part, layer, and surface properties for beams.
Part Number

Material properties and stress-free


reference temperature

Layer Number

Cross-sectional properties

Surface Number

Orientation

Beam Element Orientation


Most beams have a strong axis of bending and a weak axis of bending. Beam members are
represented as a line, and a line is an object with no inherent orientation of the cross section.
So, there must be a method of specifying the orientation of the strong or weak axis in threedimensional space. The surface number of the line controls this orientation.
More specifically, the surface number of the line creates a point in space, called the K-node.
The two ends of the beam element (the I- and J-nodes) and the K-node form a plane as shown
in the following image. The local axes define the beam elements. Axis 1 is from the I-node to
the J-node. Axis 2 lies in the plane formed by the I-, J- and K-nodes. Axis 3 is formed by the
right-hand rule. With the element axes set, the cross-sectional properties A, Sa2, Sa3, J1, I2,
I3, Z2, and Z3 can be entered appropriately in the Element Definition dialog box.

Figure 4: Axis 2 Lies in the Plane of the I-, J-, and K-nodes

For example, the following image shows part of two models, each containing a W10x45 Ibeam. Both members have the same physical orientation. The webs are parallel. However, the
analyst chose to set the K-node above the beam element in model A and to the side of the
beam element in model B. Even though the cross-sectional properties are the same, the
moment of inertia about axis 2 (I2) and the moment of inertia about axis 3 (I3) must be entered
differently.

Figure 5: Enter Cross-Sectional Properties Appropriately for Beam Orientations


Caution: Even though either properties definition scheme (Model A or Model B) is
possible, we strongly suggest that you use the method shown in the Model A
example. This method is consistent with the properties given in each of the preinstalled beam cross-section libraries. Maintaining a consistent approach
minimizes the likelihood of errors.

Default K-Node Location (for Beam Orientation by the Surface Number of


the Lines)
Table 1 shows where the K-node occurs for various surface numbers. The first choice
location is where the K-node is created provided the I-, J-, and K-nodes form a plane. If the
beam element is collinear with the K-node, then a unique plane cannot be formed. In this
case, the second choice location is used for that element.
Table 1: Correlation of Surface Number and K-Node (Axis 2 Orientation)

Surfa
1
2
3
4

Table 1: Correlation of Surface Number and K-Node (Axis 2 Orientation)

Surfa
5
6

Notice that the coordinates of the K-nodes are at an extreme distance from the origin. There is
a very good reason for this placement. Consider the limited dimensions of any real-world
analysis model. The width, length, and height are insignificant as compared to the distance to
the K-node location. For example, if +Z is the upward direction, and you use Surface 2 for
the floor joists of a 50 meter wide structure. The deviation of the vertical axis of the beams
due to them being displaced +/- 25m (or even 50m) from the Z-axis is infinitesimal. Suppose
your length unit is millimeters. The Arc Sin of (50,000mm / 1E14mm) is 0.0000000143
(which is zero for all practical purposes). So, you don't have to define multiple K-nodes for
vertically oriented beams even when they are positioned at multiple points across a very wide
structure.
You can change the surface number, hence the default orientation of the beams. Select the
appropriate beam elements (use the Selection Select Lines command) and right-click in
the display area. Choose the Edit Attributes command and change the value in the Surface:
field.
User-Defined Beam Element Orientation
In some situations, a global K-node location may not be suitable. In this case, select the beam
elements in the FEA Editor environment using the Selection Select Lines command and
right-click in the display area. Select the Beam Orientations New.. command. Type in the
X, Y, and Z coordinates of the K-node for these beams. To select a specific node in the model,
click the vertex, or enter the vertex ID in the ID field. A blue circle appears at the specified
coordinate. The following image shows an example of a beam orientation that needs the
origin defined as the k-node.

Figure 6: Skewed Beam Orientation


How to Reverse the Direction of Axis 1
The direction of axis 1 can be reversed in the FEA Editor by selecting the elements to change
(Selection Select Lines), right-clicking, and choosing Beam Orientations Invert I
and J Nodes. This ability is useful for loads that depend on the I and J nodes and for

controlling the direction of axis 3. (Recall that axis 3 is formed from the right-hand rule of
axes 1 and 2.) If any of the selected elements have a load that depends on the I/J orientation,
you choose whether or not to reverse the loads. Since the I and J nodes are being swapped,
choose Yes to reverse the input for the load and maintain the current graphical display. The I
and J nodes are inverted, and the I/J end with the load is also inverted. Choose No to keep the
original input, so an end release for node I switches to the opposite end of the element since
the position of the I node is changed.
How to Verify the Correct Axis Orientation
The orientation of the elements can be displayed in the FEA Editor environment using the
View Visibility Object Visibility Element Axis commands. The orientation can also
be checked in the Results environment using the Results Options View Element
Orientations command. Choose to show the Axis 1, Axis 2, and/or Axis 3 using red, green,
and blue arrows, respectively. See the following figure.

Figure 7: Beam Orientation Symbols (different arrows are used for each axis.)

Specify Cross-Sectional Properties of Beam Elements


The Sectional Properties table in the Cross-Section tab of the Element Definition dialog
box is used to define the cross-sectional properties for each layer in the beam element part. A
separate row appears in the table for each layer in the part. The sectional property columns
are:

Figure 8: Legend for Defining Cross-Sectional Properties

A: Specify the cross-sectional area in this column. It is the area of the


beam resisting the axial force (=FxL/(AxE)). This area must be greater
than 0.0.

J1: Specify the torsional resistance in this column. The torsional resistance
is the area moment of inertia resisting the torsional moment M1. The
angle of twist within an element is calculated by =M1xL/(J1xG) where L is
the length and G is the shear modulus. For most cross-sections, the
torsional resistance is much less than the polar moment of inertia. (For a
circular section, J1 equals the polar moment of inertia.) The torsional
resistance must be greater than 0.0.

I2: Specify the area moment of inertia about the local 2 axis in this
column. (It is also referred to as I2-2.) The local 2 axis passes through the
neutral axis of the cross section and is in the plane formed by the element
and the k-node. (See previous paragraph.). The moment of inertia must be
greater than 0.0.

I3: Specify the area moment of inertia about the local 3 axis in this
column. (It is also referred to as I3-3.) The local 3 axis passes through the
neutral axis of the cross section and forms the right-hand rule with the
element (axis 1) and axis 2. The moment of inertia must be greater than
0.0.

S2: Specify the section modulus about the local 2 axis in this column. The
section modulus is calculated from S2=I2/C3max, where C3max is
measured parallel to the 3 axis from the neutral axis to the furthermost
point on the cross section. This value is not required but is necessary for
the bending stress calculation about axis 2 (=M2/S2). If this value is 0.0,
the bending stress about the local 2 axis is set to 0.

S3: Specify the section modulus about the local 3 axis in this column. The
section modulus is calculated from S3=I3/C2max, where C2max is
measured parallel to the 2 axis from the neutral axis to the furthermost
point on the cross section. This value is not required but is necessary for
the bending stress calculation about axis 3 (=M3/S3). If this value is 0.0,
the bending stress about the local 3 axis is set to 0.

Sa2: Specify the shear area parallel to the local 2 axis. The shear area is
the effective beam cross-sectional area resisting the shear force R2 (shear
force parallel to axis 2). If the shear area is 0.0, the shear deflection in the
local 2 direction is ignored (usually a safe assumption). The shear area
correction is only needed if the beam width is comparable to the beam
length.

Sa3: Specify the shear area parallel to the local 3 axis. The shear area is
the effective beam cross-sectional area resisting the shear force R3 (shear
force parallel to axis 3). If the shear area is 0.0, the shear deflection in the
local 3 direction is ignored (usually a safe assumption). The shear area
correction is only needed if the beam width is comparable to the beam
length.
Note: Hand calculations for the deflection of beams rarely include the
effects due to shear within a beam. For example, the well-known
equations for the maximum deflection for a cantilever beam and simply
supported beam due to a point load (FL 3/(3EI) and FL 3/(48EI),
respectively) only consider the bending effects. If shear effects are
included in the finite element analysis by entering values for Sa2 and Sa3,
the calculated displacements can be higher than the hand calculations.

For standard cross-sections, or for common shapes with known dimensions, you do not have
to enter the individual properties listed above (see the next two sections on this page).
Tip: See the page Variable Cross-Section Wizard to generate a series of crosssections along the length of a beam to approximate a tapered beam.

Use Cross-Section Libraries


To use the cross-section libraries, first select the layer for which you want to define the crosssectional properties. After the layer is selected, click the Cross-Section Libraries button.
How to Select a Cross Section from an Existing Library
1. Select the library in the Section database: drop-down menu. Multiple
versions of Beam Cross-Section Libraries are provided with the software.
(Note: The libraries are set up with IYY (as defined in the AISC manual)
corresponding to I2 in the software.)

2. Select the cross section type using the Section type pull-down menu. The
type designations available for each of the AISC databases are given in
Table 5 of the Beam Cross-Section Libraries page. Additionally, you can
use the information in Table 2 of the same page to correlate shape
designations from non-AISC libraries with the corresponding AISC section
types. This table is not comprehensive because the AISC libraries do not
contain all of the cross-section types available in some of the non-AISC
libraries.
3. Select the cross section name in the Section name: list. You can search
for a name by typing a string in the field above the list.
4. Review the values in the Cross-sectional properties section of the
dialog box. If they are acceptable, click OK.
Note: To visualize the beam as a 3D solid in the Results environment, the crosssection must be chosen from the AISC 2001, AISC 2005, or AISC 14.1 (2013)
database.

Create a New Library and Add Cross-Sections to a Library


You cannot modify a pre-installed library in any way. However, you can create custom
libraries and add user-defined cross-sections to them. You can also copy a library you make at
one computer to a new computer and add it to the program. The procedure for creating
custom libraries and cross-sections is outlined on the page Create Custom Cross-Section
Libraries and Shape Entries.

How to Define Common Shapes


There is a pull-down menu in the upper right corner of the Cross-Section Libraries dialog box
that contains a list of common cross-sections. The included shapes are:

Round

Pipe

Rectangular

Hollow Rectangular

Wide-Flange Beam

C Channel

L-type (angle)

User-Defined

For all but the last type (User-Defined) you specify the shape by entering the appropriate
dimensions. A sketch and data input fields appear after you select a shape. The cross-sectional
properties (A, J1, I2, and so on) are calculated for you. Beam elements with these pre-defined
shapes can be visualized in 3D within the Results environment.

For User-Defined sections, fill in the data fields in the Cross-sectional properties section in
the middle of the dialog box. Sections defined in this manner cannot be visualized in 3D
within the Results environment.
To dimension a common shape or specify user-defined properties:
1. Go to the Element Definition dialog box for the beam part.
2. Select a layer (row) in the Sectional Properties table and click the CrossSection Libraries button.
3. Choose the User Defined option from the Section database pull-down
menu.
4. Select the desired common shape from the pull-down menu in the upperright corner of the dialog box.
o

Alternatively, for cross-sections other than the available common


shapes, enter the Cross-sectional properties in the data fields at the
center of the dialog box.

5. For all but the User Defined option, enter the appropriate dimensions
shown at the right side of the dialog box. The cross-sectional properties
are calculated and displayed.
6. Click OK.
7. Repeat steps 2 through 6 for the next layer, or click OK to exit the
Element Definition dialog box.

Other Beam Element Parameters


In addition to the cross-sectional properties, the only other parameter for beam elements is
the stress free reference temperature. It is specified in Stress Free Reference Temperature
field in the Thermal tab of the Element Definition dialog box. This value is used as the
reference temperature to calculate element-based loads associated with constraint of thermal
growth using the average of the nodal temperatures. The value you enter in the Default nodal
temperature field in the Analysis Parameters dialog box determines the global
temperatures on nodes that have no specified temperature.

Basic Steps to Use Beam Elements


1. Be sure that a unit system is defined.
2. Be sure that the model is using a structural analysis type.
3. Right-click the Element Type heading for the part that you want to be
beam elements..
4. Select the Beam option.

5. Right-click the Element Definition heading.


6. Select the Edit Element Definition command.
7. In the Cross Section tab, enter the proper cross- sectional properties for
each layer of beams. To use saved properties, common shapes, or
standard sections, press the Cross-Section Libraries button. (See the
preceding Use Cross-Section Libraries and How to Define Common Shapes
sections.
8. Once your sectional properties are entered, click OK.

A series of linearly varying beam cross-sections can be created using the Variable CrossSection Wizard. Each beam element has a different cross-section dimension based on
interpolating the two user-specified end dimensions. The wizard creates a stepped beam that
approximates a tapered beam. Any user-defined shape (rectangle, wide-flange beam, channel,
pipe, and so on) whose dimensions can be entered in the normal cross-section library dialog
box can be used for the variable cross-section.

Figure 1: Variable Cross-Section Beam - four different mesh densities are shown to
simulate a tapered beam with a variable height (bottom figure).
To generate the variable cross-sections

1. Draw the tapered beam member on a unique part number. Use the surface number
necessary for the proper orientation (see Beam Element Orientation on the page Beam
Elements), but any layer number can be used. (The layer number is by the wizard.) all
the lines in the part must connect together so that there are just two endpoints. The
lines must be continuous and not branch or create a loop. They are not required to be
in a straight line.
2. Divide the member into the number of elements needed. Since the taper is
approximated, the results are more accurate if more elements are used.
3. Set the Element Type to Beam.
4. Select the part (Selection
Section Wizard.

Select

Parts), right-click, and choose Variable Cross-

5. Select the type of cross-section to use from the drop-down.


6. Enter the dimensions of the cross-section for Vertex A and Vertex B. Note the symbols
A and B added to each end of the part in the display area. A dimension of 0 cannot be
used for some sections; enter a small value instead.
7. Click OK.
The wizard changes the model.

Each line segment of the part is changed to a unique layer number, starting with layer
1, and ending with layer n.

The cross-sectional dimensions for the n layers are interpolated at the midpoint of
each segment using the cumulative length from point A to the midpoint.

The cross-sectional properties for the interpolated dimensions are calculated and
entered into the Sectional Properties spreadsheet of the Element Definition.
Additional input may be required in the Element Definition to complete the definition
of the beam.

Note:

Once the wizard completes, the part is identical to what is created by hand if you
chose to enter the input for each element. I any manual changes are made to the part ,
such as making the member longer or shorter, rerun the wizard to compute the new
values.

Since the cross-section is calculated at the midpoint of each element, the stress at each
end of the element is either an overestimate or underestimate of the theoretical result.
Use more elements in the part to minimize the approximation.

Gap Elements

Gap elements are two-node elements formulated in three-dimensional space. This element
type is only available in a static stress analysis with linear material models.
Two end nodes specified in three-dimensional space define gap elements. Only the axial
forces of the element are calculated for each element, and depending on the settings, only
compressive forces or only tensile forces are generated. No element-based loading is defined
for gap elements.
A compression gap is not activated until the gap is closed; a tension gap is not activated until
the gap is opened. Therefore, the structural behavior of a finite element model associated with
gap elements is always nonlinear because of its indeterminate condition. Whether the gaps
are closed or opened is not known in advance. An iterative solution method is used to
determine the status (opened or closed) of the gap elements.
Since the analysis is linear and small deflection theory is used, only motion in the direction of
the original gap element orientation is considered. Sideways motion does not affect the status
of the gap element.
In general, there are three applications for gap elements. Each has its own characteristics in
terms of element input. They are briefly summarized as follows:

A
Rigid support at the structure boundary to calculate the support reactions
Interface element between two faces of the structure in space
Elastic spring between the base of the structure and the foundation

Avoid excessively stiff gap elements (with large spring stiffness) that are not aligned with the
global coordinate system. Such elements introduce large off-diagonal values into the
structural stiffness matrix and cause solution difficulties. The resulting solution may also be
inaccurate. The provided spring stiffness, about three or four orders of magnitude larger than
the other normal stiffnesses in the structure, is sufficient for rigid gap elements used in
application type (1).
Note: This content applies to gap elements created by hand (lines drawn
between two nodes of the model or between the model and the ground). Gap or
contact elements created automatically (CAD models, 2D automatic meshes, or
hand-built models) are slightly different. See the Types of Contact page.

Use Gap Elements

To model the effects of a spring or cable where the stiffness is not always
present under all loadings.

To find the contact force between two parts under a load.

Gap Element Parameters


When using gap elements, first select the type of gap element to use for the part in the Type
drop-down menu in the Element Definition dialog box. The options are:
Type
Compression with Gap
Tension with Gap
Compression without Gap
Tension without Gap

The next step is to define the stiffness of the gap elements in the Stiffness field. See the table
in the previous paragraph, What is a Gap Element, for guidelines on the stiffness.
When duplicating a real spring (tension or compression) or chain-like arrangement (tension
only), enter the known stiffness. The stiffness (k) of a rod or simple wire can be calculated
from k=A*E/L, where A is the cross-sectional area, E is the modulus of elasticity, and L is the
length of the rod. When duplicating part-to-part contact, a rigid stiffness is required. A
stiffness on the same order of magnitude as the modulus of the material is sufficient. Even
when the two values are in different units (force/length versus force/length squared). Another
method of calculating the stiffness is to use the definition of stiffness: k = F/ where F is the
force transmitted through the element and is the compression or elongation in the element.
Based on the model, a reasonable can be chosen. If the contact force can be estimated, the
required stiffness can be calculated.

To Use Gap Elements


1. Be sure that a unit system is defined.
2. Be sure that the model is using the static stress with linear material
models analysis type.
3. Draw the gap elements as lines. See Tips for Drawing Gap Elements below.
4. Right-click the Element Type heading for the part that you want to be
gap elements.
5. Select the Gap command.
6. Right-click the Element Definition heading.
7. Select the Edit Element Definition heading.
8. In the Type drop-down box select the type of gap element that you want
to apply.

If the element should be active only when the new length is less
than zero, select the Compression with Gap option.

If the element should be active only when the new length is greater
than twice the original length, select the Tension with Gap option.

If the element should be active only when the new length is less
than the original length, select the Compression without Gap
option.

If the element should be active only when the new length is greater
than the original length, select the Tension without Gap option.

9. Enter the stiffness of the gap element in the Stiffness field. This is
required information and the processor will not run without a value.
10.Press the OK button.
11.Add boundary elements with small stiffness as necessary to stabilize the
parts. See Perform Analyses with Gap Elements for additional information.

Tips for Drawing Gap Elements


1. Draw lines connecting the corresponding nodes. Use the Draw
Line command.

Draw

2. Use the Draw


Design
Contact Elements command. This command
automatically creates the lines between two sets of vertices. In the case of
linear stress (where small deformation is assumed), only the shortest line
perpendicular between the faces is normally required. In this situation, use
the Constrain lengths option. See the paragraph Add Contact Elements to
a Model on the Add Geometry page.
3. Copy the lines from one face to the corresponding face, and in the process
use the option to join the copies. Then define the joined lines as gap
elements. Imagine two plates in contact. Instead of drawing the lines for
the gap elements one by one, do the following:
a. Select the lines (Selection
faces.

Select

b. Start the copy command with Draw

Lines) on one of the matching


Pattern

Move or Copy.

c. Activate the check box Copy. Make sure the number of copies to
make is set to one.
d. Activate the check box Join.
e. Set the total distance and direction vector for the copy. Use the
Vector Selector button to click two vertices if you are not sure of the
distance or direction.

f.

Click OK. The selected lines are copied to the matching face and
lines between the faces that become the gap elements.

g. Since this creates a copy of the original lines on the matching face,
and since these extra lines are not needed, click Delete. (The last
copy made is still selected after the command completes.)
h. Last step: change the join lines to a part number that can be
defined as gap elements. This step typically consists of using a box
selection (Selection Shape Rectangle) to select the joined lines,
then right-click, Edit Attributes and enter an appropriate part
number. If using a box selection, the entire line must be inside the
selection rectangle. Consequently, the lines on the faces of the
parts are also selected. Do not change the attributes of those lines.
Unselect the lines on the faces by using the subtract mode (hold the
Shift and Ctrl keys) while box selecting those lines. (See the Modify
Attributes and Duplicate page.)

See Perform Analyses with Gap Elements for additional information common to gap and
surface contact elements.
https://knowledge.autodesk.com/support/simulation-mechanical/learnexplore/caas/CloudHelp/cloudhelp/2016/ENU/SimMech-UsersGuide/files/GUIDF9B8B578-19D6-4FDB-8EC5-BE3FEFEB7C22-htm.html

You might also like