0% found this document useful (0 votes)
56 views

Reference

The document discusses material properties and databases in PLAXIS 2D. There are seven types of material sets for soils, discontinuities, plates, geogrids, embedded beams, cable bolts, and anchors that are stored in the material database. The material database can be accessed to assign properties to objects in the geometry model. It provides details on the various material models that can be selected and the parameters involved for modeling soil and interface behavior.

Uploaded by

thach
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
56 views

Reference

The document discusses material properties and databases in PLAXIS 2D. There are seven types of material sets for soils, discontinuities, plates, geogrids, embedded beams, cable bolts, and anchors that are stored in the material database. The material database can be accessed to assign properties to objects in the geometry model. It provides details on the various material models that can be selected and the parameters involved for modeling soil and interface behavior.

Uploaded by

thach
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 139

Material properties and material database

6
In PLAXIS 2D, soil properties and material properties of structures are stored in material data sets. There are
seven different types of material sets grouped as data sets for:
• Soil and interfaces
• Discontinuities
• Plates
• Geogrids
• Embedded beams
• Cable bolts
• Anchors
All data sets are stored in the material database. From the database, the data sets can be assigned to the soil
clusters or to the corresponding structural objects in the geometry model.
The material database can be opened by:
1. Selecting the option Show materials in the Soil menu of the Soil mode or in the Structures menu of the
Structures mode.
2.
Clicking on the Show materials button available in the Soil, Structures and Staged construction modes.
As a result, the Material sets window appears showing:
1. The contents of the Project materials database.
2. The window can be expanded to show the Global materials database by clicking the Show global button in
the upper part of the window.
The Material sets window displaying the material defined in the current project and the ones available in a
selected global database is shown in Figure 116 (on page 163).

PLAXIS 162 PLAXIS 2D-Reference Manual


Material properties and material database

Figure 116: Material sets window showing the project and the global database

The database of a new project is empty. The global database can be used to store material data sets in a global
folder and to exchange data sets between different projects.
At both sides of the window (Project materials and Global materials) there are two drop-down menus and a
tree view. The Set type can be selected from the drop-down menu on the left hand side. The Set type parameter
determines which type of material data set is displayed in the tree view (Soil and interfaces, Discontinuities,
Plates, Geogrids, Embedded beams, Cable bolts and Anchors).
The data sets in the tree view are identified by a user-defined name. The data sets for Soil and interfaces can be
ordered in groups according to the material model, the material type or the name of the data set by selecting this
order in the Group order drop-down menu. The None option can be used to discard the group ordering.
The small buttons between the two tree views can be used to copy individual data sets from the project database
to the selected global database or vice versa.
• To copy the selected project material set to the global database.
• To copy all the project material sets of the specified type to the global database.
• To copy the selected global material set to the project database.
The location of the selected global database is shown below its tree view. The buttons below the tree view of the
global database enable actions in the global database.

PLAXIS 163 PLAXIS 2D-Reference Manual


Material properties and material database

Select To select an existing global database.

Delete To delete a selected material data set from the selected global database.

By default, created data sets in the Global databases materials are stored in a location that can be defined by
the user and in different files as follows:
• Soil and interface materials will be contained in a file named by default as Soil.matXdb'. This file is
compatible with other PLAXIS 2D database files for Soil and interface.
• Material data sets for structural elements such as Discontinuities, Plates, Geogrids, Embedded beams,
Cable bolts and Anchors will be contained in separate files. Respectively, the default name for these
materials will be 'Discontinuity.matXdb', 'Plate.matXdb', 'Geogrid.matXdb', 'EmbeddedBeam.matXdb',
'CableBolt.matXdb', and 'Anchor.matXdb'.

Note:
• A new global database can be created by clicking the Select button, defining the name of the new global
database and clicking Open.

The project data base can be managed using the buttons below the tree view of the project database.

To create a new data set in the project. As a result, a new window appears in which the
material properties or model parameters can be entered. The first item to be entered is
New always the Identification, which is the user-defined name of the data set. After
completing a data set, it will appear in the tree view, indicated by its name as defined by
the Identification.

Edit To modify the selected data set in the project material database.

To perform standard soil lab tests. A separate window will open where several basic soil
tests can be simulated and the behaviour of the selected soil material model with the
SoilTest
given material parameters can be checked ( Simulation of soil lab tests [GSE] (on page
498)).

Copy To create a copy of a selected data set in the project material database.

Delete To delete a selected material data set from the project material database.

Discard To revert the changes and fire a command when clicked.

Note:
In a Material dataset window (Material properties and material database (on page 162)) it is possible to go
through the options with arrows and/or Enter. When a property such as Material model or Drainage type, is
reached, the drop-down menu can be activated by pressing the Space key. Arrows and /or letters can be used to
make a selection, that is finalised by pressing Enter.

Each Soil and structural element material of the database contains the parameters characterizing the material.
To prevent an invalid definition of the material properties a Feedback side panel is included as displayed in
Figure 117 (on page 165). Three types of feedback messages are possible:

PLAXIS 164 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

• Errors: the parameter value or combination of parameter values must be changed, otherwise the material
set could be invalid and calculation of the project will be blocked.
• Warnings: the parameter value seems to deviate from a recommended parameter value or parameter range.
Generally the material set will not be considered invalid and calculating the project will not be blocked. The
chosen parameter could however cause unexpected results.
• Hints: the entered parameter can be defined under certain circumstances or options.

Figure 117: Feedback side panel - Example of feedback messages

Note: Once the all errors, warnings and hints are attended the Feedback side panel disappears.

6.1 Modelling soil and interface behaviour


The material properties and model parameters for soil clusters are entered in material data sets Figure 118 (on
page 166). The properties in the data sets are divided into six tabsheets: General, Mechanical, Groundwater,
Thermal, Interfaces and Initial.

6.1.1 General tabsheet


The General tabsheet contains the type of soil model, the drainage type and the general soil properties such as
unit weights. Several data sets may be created to distinguish between different soil layers. A user may specify

PLAXIS 165 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

any identification title for a data set in the General tabsheet of the Soil window. It is advisable to use a
meaningful name since the data set will appear in the database tree view by its identification.
For easy recognition in the model, a colour is given to a certain data set. This colour also appears in the database
tree view. PLAXIS 2D selects a unique default colour for a data set, but this colour may be changed by the user.
Changing the colour can be done by clicking on the colour box in the General tabsheet.

Figure 118: General tabsheet of the Soil window

6.1.1.1 Material model


Soil and rock tend to behave in a highly non-linear way under load. This non-linear stress-strain behaviour can
be modelled at several levels of sophistication. Clearly, the number of model parameters increases with the level
of sophistication. PLAXIS 2D supports different models to simulate the behaviour of soil and other continua. The
models and their parameters are described in detail in the Material Models Manual. A short discussion of the
available models is given below:
Linear elastic model: This model represents Hooke's law of isotropic linear elasticity. The linear elastic model
is too limited for the simulation of soil behaviour. It is primarily used for stiff structures in the soil.
Mohr-Coulomb model (MC): This well-known linear elastic perfectly-plastic model is used as a first
approximation of soil behaviour in general. It is recommended to use this model for a first analysis of the
problem considered. A constant average stiffness is estimated for the soil layer. Due to this constant stiffness,
computations tend to be relatively fast and a first estimate of deformations can be obtained.
Hardening Soil model (HS): This is an advanced model for the simulation of soil behaviour. The Hardening Soil
model is an elastoplastic type of hyperbolic model, formulated in the framework of shear hardening plasticity.
Moreover, the model involves compression hardening to simulate irreversible compaction of soil under primary

PLAXIS 166 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

compression. This second-order model can be used to simulate the behaviour of sands and gravel as well as
softer types of soil such as clays and silts.
Hardening Soil model with small-strain stiffness (HSsmall): This is an elastoplastic type of hyperbolic model,
similar to the Hardening Soil model. Moreover, this model incorporates strain dependent stiffness moduli,
simulating the different reaction of soils from small strains (for example vibrations with strain levels below 10-5)
to large strains (engineering strain levels above 10-3).
Modified Cam-Clay model (MCC): This well-known critical state model can be used to simulate the behaviour
of normally consolidated soft soils. The model assumes a logarithmic relationship between the void ratio and the
mean effective stress.
Hoek-Brown model (HB): This well-known elastic perfectly-plastic model is used to simulate the isotropic
behaviour of rock. A constant stiffness is used for the rock mass. Shear failure and tension failure are described
by a non-linear strength curve.
Jointed Rock model (JR): This is an anisotropic elastic-perfectly plastic model where plastic shearing can only
occur in a limited number of shearing directions. This model can be used to simulate the anisotropic behaviour
of stratified or jointed rock.
NGI-ADP model (NGI-ADP): The NGI-ADP model may be used for capacity, deformation and soil-structure
interaction analysis involving undrained loading of clay-type materials. Distinct anisotropic shear strengths may
be defined for different stress paths.
[ADV] Soft soil model (SS): This is a Cam-Clay type model that can be used to simulate the behaviour of soft
soils like normally consolidated clays and peat. The model performs best in situations of primary compression.
[ADV] Soft Soil Creep model (SSC): This is a second order model formulated in the framework of
viscoplasticity. The model can be used to simulate the time-dependent behaviour of soft soils like normally
consolidated clays and peat. The model includes logarithmic primary and secondary compression.
[ADV] UDCAM-S : The UDCAM-S model is a derived NGI-ADP model to deal with undrained soil behaviour and
degradation of the strength and stiffness in cyclic loading of clay or very low permeable silty soils. It implements
a pre-processing procedure called Cyclic accumulation and optimisation tool ( Cyclic accumulation and
optimisation tool [ADV] (on page 213)) to obtain the degraded parameter set based on the type of analysis.
[ADV] Sekiguchi-Ohta model (Inviscid): The Sekiguchi-Ohta model (Inviscid) is a Cam-Clay type effective
stress model for time-independent behaviour of clay-type soils.
[ADV] Sekiguchi-Ohta model (Viscid): The Sekiguchi-Ohta model (Viscid) is a Cam-Clay type effective stress
model for time-dependent behaviour (creep) behaviour of clay-type soils.
[ADV] Concrete model: The Concrete modelis an advanced elastoplastic model for concrete and shotcrete
structures. It simulates the time-dependent strength and stiffness of concrete, strain hardening-softening in
compression and tension as well as creep and shrinkage. The failure criterion involves a Mohr-Coulomb yield
surface for deviatoric loading, which is combined with a Rankine yield surface in the tensile regime.
[ULT] UBC3D-PLM : This is an advanced model for the simulation of liquefaction behaviour in dynamics
applications. The model includes accumulation of irreversible strains during cyclic loading. In combination with
undrained behaviour it accumulates pore pressures, which may eventually lead to liquefaction.
[ADV]/[ULT]+[GSE] User-defined Soil Models (UDSM): With this option it is possible to use other constitutive
models than the standard PLAXIS 2D models. For a detailed description of this facility, reference is made to the
Material Models Manual. Links to existing UDSM as well as all models are available on the PLAXIS Knowledge
Base.

PLAXIS 167 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

6.1.1.2 Drainage type


In principle, all model parameters in PLAXIS 2D are meant to represent the effective soil response, i.e. the
relationship between the stresses and the strains associated with the soil skeleton. An important feature of soil
is the presence of pore water. Pore pressures significantly influence the (time-dependent) soil response. PLAXIS
2D offers several options to enable incorporation of the water-skeleton interaction in the soil response. The
most advanced option is a Fully coupled flow-deformation analysis. However, in many cases it is sufficient to
analyse either the long-term (drained) response or the short-term (undrained) response without considering
the time-dependent development of pore pressures. In the latter case (undrained), excess pore pressures are
generated as a result of stress changes (loading or unloading). The dissipation of these excess pore pressures
with time can be analysed in a Consolidation calculation.
The simplified water-skeleton interaction, as considered in a Plastic calculation, a Safety analysis or a Dynamics
analysis, is defined by the Drainage type parameter. PLAXIS 2D offers a choice of different types of drainage:
Drained behaviour: Using this setting no excess pore pressures are generated. This is clearly the case for dry
soils and also for full drainage due to a high permeability (sands) and/or a low rate of loading. This option may
also be used to simulate long-term soil behaviour without the need to model the precise history of undrained
loading and consolidation.
Undrained behaviour: This setting is used for saturated soils in cases where pore water cannot freely flow
through the soil skeleton. Flow of pore water can sometimes be neglected due to a low permeability (clays)
and/or a high rate of loading. All clusters that are specified as undrained will indeed behave undrained, even if
the cluster or a part of the cluster is located above the phreatic level.
Distinction is made between three different methods of modelling undrained soil behaviour. Method A is an
undrained effective stress analysis with effective stiffness as well as effective strength parameters. This method
will give a prediction of the pore pressures and the analysis can be followed by a consolidation analysis. The
undrained shear strength (su) is a consequence of the model rather than an input parameter. It is recommended
to check this shear strength with known data. To consider this type of analysis, the Undrained A option should be
selected in the Drainage type drop-down menu.
Method B is an undrained effective stress analysis with effective stiffness parameters and undrained strength
parameters. The undrained shear strength su is an input parameter. This method will give a prediction of pore
pressures. However, when followed by a consolidation analysis, the undrained shear strength (su) is not
updated, since this is an input parameter. To consider this type of analysis, the Undrained B option should be
selected in the Drainage type drop-down menu.
Method C is an undrained total stress analysis with all parameters undrained. This method will not give a
prediction of pore pressures. Therefore it is not useful to perform a consolidation analysis. The undrained shear
strength (su) is an input parameter. To consider this type of analysis, the Undrained C option should be selected
in the Drainage type drop-down menu.
More information about modelling undrained behaviour can be found in Modelling undrained behaviour (on
page 260) and the Material Models Manual.
Non-porous behaviour: Using this setting neither initial nor excess pore pressures will be taken into account in
clusters of this type. Applications may be found in the modelling of concrete or structural behaviour. Non-porous
behaviour is often used in combination with the Linear elastic model. The input of a saturated weight is not
relevant for non-porous materials or intact rock.
In a consolidation analysis or a fully coupled flow-deformation analysis it is the permeability parameter in the
Flow tabsheet that determines the drainage capacity of a layer rather than the drainage type. Still, the drainage
type has influence on the applied compressibility of water in a consolidation analysis or a fully coupled flow-
deformation analysis. For more information see Appendix C.

PLAXIS 168 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Note:
The Drainage type setting is only considered in a Plastic calculation, a Safety analysis or a Dynamics analysis.
When a Consolidation analysis, a Fully coupled flow-deformation analysis or a Dynamics with Consolidation
analysis is performed, the Drainage type is ignored and the soil response is determined by the Permeability of the
material.

6.1.1.3 Unsaturated and saturated weight (γunsat and γsat)


The unsaturated and the saturated weights, entered as a force per unit volume, refer to the total unit weight of
the soil skeleton including the fluid in the pores. The unsaturated weight γunsat applies to all material above the
phreatic level and the saturated unit weight γsat applies to all material below the phreatic level, where the
phreatic level itself is generally defined as the level where the steady-state pore pressure is zero (psteady = 0).
Only in the case of a fully coupled flow-deformation analysis, the phreatic level is defined as the level where the
current pore water pressure is zero (pwater = 0). This means that during a fully coupled flow-deformation
analysis the position of the phreatic level and hence the material weight can change.
For non-porous material only the unsaturated weight is relevant, which is just the total unit weight. For porous
soils the unsaturated weight is obviously smaller than the saturated weight. For sands, for example, the
saturated weight is generally around 20 kN/m3 whereas the unsaturated weight can be significantly lower,
depending on the degree of saturation.
Note that soils in practical situations are never completely dry. Hence, it is advisable not to enter the fully dry
unit weight for γunsat. For example, clays above the phreatic level may be almost fully saturated due to capillary
action. Other zones above the phreatic level may be partially saturated. PLAXIS 2D can deal with partially
saturated soil behaviour above the phreatic level. However, the unit weight of soil is always defined here by
γunsat, irrespective of the degree-of-saturation.
Weights are activated by means of Gravity loading or K0 procedure in the Calculation mode, which is always the
first calculation phase (Initial phase) (see Initial stress generation (on page 311)).

6.1.1.4 Void ratio and porosity


The void ratio, e, is related to the porosity, n (e = n / (1-n)). This quantity is used in some special options. The
initial value einit is the value in the initial situation. The actual void ratio is calculated in each calculation step
from the initial value and the volumetric strain Δεv. These parameters are used to calculate the change of
permeability when input is given for the ck value (in the Flow tabsheet).

6.1.1.5 Rayleigh damping [ULT]


Material damping in dynamics calculations is caused by the viscous properties of soil, friction and the
development of irreversible strains. All plasticity models in PLAXIS 2D can generate irreversible (plastic) strains,
and may cause material damping. However, this damping is generally not enough to model the damping
characteristics of real soils. For example, most soil models show pure elastic behaviour upon unloading and
reloading which does not lead to damping at all. Some models in PLAXIS 2D include viscous behaviour, for
instance the Soft Soil Creep model. Using this model in dynamics calculations may lead to viscous damping, but
also the Soft Soil Creep model hardly shows any creep strain in load / reload cycles. There are also models in
PLAXIS 2D that include hysteretic behaviour in loading / reload cycles, like the Hardening Soil model with small-
strain stiffness ( Material Models Manual - The Hardening Soil model with small-strain). When using this model,
the amount of damping that is obtained depends on the amplitude of the strain cycles. Considering very small

PLAXIS 169 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

vibrations, even the Hardening Soil model with small-strain stiffness does not show material damping as well as
numerical damping, whereas real soils still show a bit of viscous damping. Hence, additional damping is needed
to model realistic damping characteristics of soils in dynamics calculations. This can be done by means of
Rayleigh damping.
Rayleigh damping is a numerical feature in which a damping matrix C is composed by adding a portion of the
mass matrix M and a portion of the stiffness matrix K:
C = α M +β K Eq. [7]
The parameters α and β are the Rayleigh coefficients and can be specified in the corresponding cells in the
General tabsheet of the Soil window shown in Figure 117 (on page 165).
α is the parameter that determines the influence of mass in the damping of the system. The higher α is, the more
the lower frequencies are damped. β is the parameter that determines the influence of stiffness in the damping
of the system. The higher β is, the more the higher frequencies are damped. In PLAXIS 2D, these parameters can
be specified for each material data set for soil and interfaces as well as for material data sets for plates. In this
way, the (viscous) damping characteristics can be specified for each individual material in the finite element
model.
Despite the considerable amount of research work in the field of dynamics, little has been achieved yet for the
development of a commonly accepted procedure for damping parameter identification. Instead, for engineering
purposes, some measures are made to account for material damping. A commonly used engineering parameter
is the damping ratio ξ. The damping ratio is defined as ξ = 1 for critical damping, i.e. exactly the amount of
damping needed to let a single degree-of-freedom system that is released from an initial excitation u0, smoothly
stop without rebouncing.
Considering Rayleigh damping, a relationship can be established between the damping ratio ξ and the Rayleigh
damping parameters α and β:

α + βω 2 = 2ωξ and ω = 2πf Eq. [8]


where ω is the angular frequency in rad/s and f is the frequency in Hz (1/s).
u
Overdamped(ξ>1)

Critically damped(ξ=1)

Underdamped(ξ<1)

Figure 119: Role of damping ratio ξ in free vibration of a single degree-of-freedom system

Solving this equation for two different target frequencies and corresponding target damping ratios gives the
required Rayleigh damping coefficients:
ω1ξ2 − ω2ξ1 ω1ξ1 − ω2ξ2
α = 2ω1ω2 and β = 2 2 2 Eq. [9]
ω12 − ω22 ω1 − ω2

PLAXIS 170 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

For example, when it is desired to have a target damping of 8% at the target frequencies f = 1.5 Hz and 8.0 Hz, the
corresponding Rayleigh damping ratios are α = 1.2698 and β = 0.002681. From Figure 120 (on page 171) it can
be seen that within the range of frequencies as defined by the target frequencies the damping is less than the
target damping, whereas outside this range the damping is more than the target damping.

Figure 120: Rayleigh damping parameter influence

In order to calibrate the frequencies corresponding to Target 1 and Target 2, different procedures can be found
in literature. In particular, Hudson, Idriss & Beirkae (1994) (on page 553) and Hashash & Park (2002) (on page
553) suggest to select the first target frequency as the first natural frequency of the soil deposit f1, while the
second target frequency is the closest odd integer larger than the ratio fp/f1, i.e. the predominant frequency of
the input motion (that can be determined from the input Fourier spectrum) over the natural frequency of the
soil. The natural frequency of the soil deposit of thickness H is related to its geometry and stiffness according to
the following equation:
Vs
f1 = 4H
Eq. [10]

where vs is the shear wave velocity in the soil deposit, that is a function of the shear stiffness modulus G.
Amorosi, Boldini & Ellia (2010) (on page 552) suggest to consider the frequency interval characterised by the
highest energy content that can be evaluated by plotting the Fourier spectrum at different depths of the soil
deposit and the amplification function between the surface and the base level. It has been demonstrated that this
procedure overcomes the errors that can occur with the previous procedure for increasing values of the ratio
fp/f1 and of the soil deposit thickness.
More generally, the two frequencies are identified through an iterative procedure.
It is suggested not to use the simplified Rayleigh formulation, i.e. the small strain viscous damping effects are
assumed to be proportional only to the stiffness of the soil deposit:
C = β K (not recommended) Eq. [11]
where β = 2 ξ/ω1 and ω1 is the angular frequency of the first natural mode of the soil column. It has been noticed
that the simplified Rayleigh formulation may lead to an underestimation of the site response, especially when

PLAXIS 171 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

the natural frequency of the soil deposit and the predominant frequency of the input motion are far from each
other, resulting in overestimation of damping in the high-frequency range.
When the Rayleigh damping Input method is set to SDOF equivalent, the damping parameters (α and β) will be
automatically calculated by the program when the target damping ratio (ξ) and the target frequencies (f) are
specified. The panel displayed in the General tabsheet when one of the cells corresponding to the damping
parameters is clicked. Figure 121 (on page 172) shows the damping ratio as a function of the frequency.
Alternatively, if the Input method is set to Direct the damping parameters α and β can be directly specified by
the user.

Figure 121: Input of ξ and f

6.1.2 Mechanical tabsheet


The Mechanical tabsheet contains the stiffness and strength parameters of the selected soil model. These
parameters depend on the selected soil model as well as on the selected drainage type.

PLAXIS 172 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

6.1.2.1 Linear Elastic model (LE)


The Mechanical tabsheet for the Linear Elastic model(drained behaviour) is shown in Figure 122 (on page
174).
The model involves two elastic stiffness parameters, namely the effective Young's modulus E' and the effective
Poisson's ratio γ'.

E'ref Effective Young's modulus [kN/m2]

ν' Effective Poisson's ratio [-]

During the input for the Linear Elastic model the values of the shear modulus G and the oedometer modulus Eoed
are presented as auxiliary parameters (alternatives).

Gref Shear modulus, where G = E'/(2(1 + ν' ) ) [kN/m2]

Eoed Oedometer modulus, where Eoed = E'(1-ν')/((1+ν')(1 - 2ν')) [kN/m2]

Note:
• Optional drainage types when the Linear Elastic model is selected are: Drained, Undrained A, Undrained C,
and Non-porous.
• In the case of Undrained A or Non-porous drainage types, the same parameters are used as for drained
behaviour.
• In the case of Undrained C drainage type, an undrained Young's modulus (Eu) and undrained Poisson's ratio
(νu) are used.

PLAXIS 173 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 122: Mechanical tabsheet for the Linear Elastic model(drained behaviour)

Note that the alternatives are influenced by the input values of E' and ν'. Entering a particular value for one of
the alternatives Gref or Eoed results in a change of the Young's modulus E'ref.
It is possible for the Linear Elastic model to specify a stiffness that varies linearly with depth. Therefore, the
increment of stiffness per unit of depth, E'inc, can be defined. Together with the input of E'inc the input of yref
becomes relevant. For any y-coordinate above yref} the stiffness is equal to E'ref. For any y-coordinate below yref
the stiffness is given by:
E ′ ( y ) = E ′ ref + ( yref − y )E ′ inc ; y < yref Eq. [12]
The Linear Elastic model is usually inappropriate to model the highly non-linear behaviour of soil, but it is of
interest to simulate structural behaviour, such as thick concrete walls or plates, for which strength properties
are usually very high compared with those of soil. For these applications, the Linear Elastic model will often be
selected together with Non-porous type of material behaviour in order to exclude pore pressures from these
structural elements.

Note:
When embedded beams penetrate a polygon cluster with linear elastic material behaviour, the specified value of
the shaft resistance is ignored. The reason for this is that the linear elastic material is not supposed to be soil, but
part of the structure. The connection between the pile and the structure is supposed to be rigid to avoid, for
example, punching of piles through a concrete deck.

[ULT] As a second alternative for the stiffness parameters of the soil, the velocities of wave propagation in soil
can be defined in the Mechanical tabsheet of the Soil window when the Dynamics module of the program is
available. These velocities are:

PLAXIS 174 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Vs Shear wave velocity, where V s = G0 / ρ [m/s]

Vp Compression wave velocity, where V p = E/ρ [m/s]

where ρ = γ / g.
As the wave velocities are influenced by the input values of E' and ν', entering a particular value for one of the
wave velocities results in a change of the Young's modulus.

Note:
• Velocities of wave propagation in soil can be defined only for models with stress independent stiffness.

6.1.2.2 Mohr-Coulomb model (MC)


The linear-elastic perfectly-plastic model with Mohr-Coulomb failure contour (in short the Mohr-Coulomb
model) requires a total of five parameters (two stiffness parameters and three strength parameters), which are
generally familiar to most geotechnical engineers and which can be obtained from basic tests on soil samples.

PLAXIS 175 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 123: Mechanical tabsheet for the Mohr-Coulomb model (drained behaviour)

The stiffness parameters of the Mohr-Coulomb model (drained behaviour) are:

E'ref Effective Young's modulus [kN/m2]

ν Effective Poisson's ratio [-]

Note:
• Optional drainage types when Mohr-Coulomb model is selected are: Drained, Undrained A, Undrained B,
Undrained C, and Non-porous.
• In the case of Undrained A or Non-porous drainage types, the same parameters are used as for drained
behaviour.
• In the case of Undrained B drainage type, φ = φu = 0, ψ = 0 and the undrained shear strength su is used instead
of the effective cohesion (c').

PLAXIS 176 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

• In the case of Undrained C drainage type all parameters are undrained. i.e. Eu, νu and su as undrained Young's
modulus, undrained Poisson's ratio and undrained shear strength respectively, and φ = ψ = 0.

Instead of using the Young's modulus as a stiffness parameter, alternative stiffness parameters can be entered.
These parameters, the relations and their standard units are listed below:

Gref Shear modulus, where Gref = E'ref/(2(1 + ν )) [kN/m2]

Eoed Oedometer modulus, where Eoed = E'ref(1-ν)/((1+ν)(1 - 2ν)) [kN/m2]

Note that the alternatives are influenced by the input values of E'ref and ν. Entering a particular value for one of
the alternatives Gref or Eoed results in a change of the Young's modulus E'ref.
Stiffness varying with depth can be defined in Mohr-Coulomb model by entering a value for E'inc which is the
increment of stiffness per unit of depth. Together with the input of E'inc the input of yref becomes relevant. For
any y-coordinate above yref} the stiffness is equal to E'ref. For any y-coordinate below yref the stiffness is given by:
E ′ ( y ) = E ′ ref + ( yref − y ) E ′ inc ; y < yref Eq. [13]
The strength parameters for the Mohr-Coulomb model are:

c'ref Effective cohesion [kN/m2]

φ' Effective friction angle [°]

ψ Dilatancy angle [°]

A cohesion varying with depth can be defined in Mohr-Coulomb model by entering a value for c'inc which is the
increment of effective cohesion per unit of depth. Together with the input of c'inc the input of yref becomes
relevant. For any y-coordinate above yref the cohesion is equal to c'ref. For any y-coordinate below yref the
cohesion is given by:
c ′ ( y ) = c ′ ref + ( yref − y )c ′ inc ; y < yref Eq. [14]
In some practical problems an area with tensile stresses may develop. This is allowed when the shear stress is
sufficiently small. However, the soil surface near a trench in clay sometimes shows tensile cracks. This indicates
that soil may also fail in tension instead of in shear. Such behaviour can be included in a PLAXIS 2D analysis by
selecting the Tension cut-off option. When selecting the Tension cut-off option the allowable tensile strength may
be entered. The default value of the tensile strength is zero.
[ULT] As a second alternative for the stiffness parameters, the velocities of wave propagation in soil can be
defined in the Mechanical tabsheet of the Soil window. These velocities are:

Vs Shear wave velocity, where V s = G/ρ [m/s]

Vp Compression wave velocity, where V P = E/ρ [m/s]

where ρ = ν / g.
The wave velocities are influenced by the input values of E' and ν', hence entering a particular value for one of
the wave velocities results in a change of the Young's modulus.

PLAXIS 177 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Note:
• Velocities of wave propagation in soil can be defined only for models with stress independent stiffness.

6.1.2.3 Hoek-Brown model (HB)


The Mechanical tabsheet for the Hoek-Brown model is shown in Figure 124 (on page 178). The Hoek-Brown
parameters can be obtained from the tools available in the right-hand panel.

Note:
• Optional drainage types when Hoek-Brown model is selected are: Drained and Non-porous.
• The same parameters are used for both Non-porous drainage type and for drained behaviour.

Figure 124: Mechanical tabsheet for the Hoek-Brown model (drained behaviour)

The stiffness parameters of the Hoek-Brown model are:

PLAXIS 178 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Erm Rock mass Young's modulus [kN/m2]

ν Poisson's ratio [-]

Gref Rock mass shear modulus, Gref = Erm/(2(1+ν)) [kN/m2]

Where Gref is an alternative input for the rock mass Young's modulus Erm.
The strength parameters have been divided in 4 subcategories: the Uniaxial compressive strength, the Hoek-
Brown parameters, the Rock mass parameters and the Dilatancy.

|σci| Uni-axial compressive strength of the intact rock (>0) [kN/m2]

For the Hoek-Brown and Rock mass parameters the user has the choice how to enter them by setting the
Determination option. When this option is set to Derived, the user is asked to specify these well-known
engineering parameters:

mi Material constant for the intact rock [-]

GSI Geological Strength Index [-]

D Disturbance factor which depends on the degree of disturbance to which the rock [-]
mass has been subjected

When any of these parameters is selected, a side panel opens to supports the selection of the parameters.
However, when Determination is set to Direct it becomes possible to specify the Hoek-Brown parameters
directly:

mb Reduced intact rock parameter [-]

s Hoek-Brown material parameter [-]

a Hoek-Brown power parameter [-]

σt Tensile strength of the rock mass [kN/m2]

σc Compressive strength of the rock mass [kN/m2]

The Hoek-Brown model simulates the Tensile strength of the rock mass via the |σci parameter, calculated from
rock properties σc, s and mb (see Material Models Manual for more information on model parameters). To
enhance modelling capabilities and flexibility, PLAXIS 2D allows to calibrate separately the rock tensile strength
using tension cut-off facility. By default the tension cut-off option is disabled. When enabled, users can put a
tensile strength value, and if that value is lower than σt, the tensile capacity will be cut-off at that value. This
feature can be helpful in some situations, for instance in Safety analysis where this tension cut-off value (but not
theσt parameter) is reduced with the safety factor.
Finally, the dilatancy can be specified through the following parameters:

ψmax Dilatancy at zero stress level [°]

PLAXIS 179 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

σψ Stress level at which dilatancy is fully suppressed [kN/m2]

6.1.2.4 Jointed Rock model (JR)


The Mechanical tabsheet for the Jointed Rock model is shown in Figure 125 (on page 180).

Note:
• Optional drainage types when Jointed Rock model is selected are: Drained and Non-porous.
• The same parameters are use for both thef Non-porous drainage type and drained behaviour.

Figure 125: Mechanical tabsheet for the Jointed Rock model

Parameters for stiffness:

Et Young's modulus for rock as a continuum, i.e. in 'Plane 1' direction [kN/m2]

νnt Poisson's ratio for rock as a continuum, i.e. in 'Plane 1' direction [-]

PLAXIS 180 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Anisotropic elastic parameters 'Plane 1' direction (e.g. stratification direction):

En Young's modulus perpendicular to 'Plane 1' direction [kN/m2]

νts Poisson's ratio perpendicular to 'Plane 1' direction [-]

Gnt Shear modulus perpendicular to 'Plane 1' direction [kN/m2]

Parameters for strength:


Strength parameters in joint directions (Plane i=1, 2, 3):

ci Cohesion [kN/m2]

φi Friction angle [°]

ψi Dilatancy angle [°]

σt,i Tensile strength [kN/m2]

Definition of joint directions (Plane i=1, 2, 3):

Number of Number of joint directions (1≤n≤3) [-]


planes

α1,i Dip angle (visualized in the side panel) [°]

α2,i Dip direction [°]

Weak planes (joints) may fail in shear (modeled by Coulomb strength parameters cohesion and friction angle) or
in tension. The latter can be included in a PLAXIS 2D analysis by selecting the Tension cut-off option. When
selecting the Tension cut-off option the allowable tensile strength may be entered. The default value of the tensile
strength is zero.

6.1.2.5 Hardening Soil model (HS)


The Mechanical tabsheet for the Hardening Soil modelis shown in Figure 126 (on page 182).

Note:
• Optional drainage types when Hardening Soil model is selected are: Drained, Undrained A, and Undrained B.
• In the case of Undrained A drainage type, the same parameters are used as for drained behaviour.
• In the case of Undrained B drainage type, φ = φu = 0 , ψ = 0 and the undrained shear strength su is used
instead of the effective cohesion (c').

PLAXIS 181 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 126: Mechanical tabsheet for the Hardening Soil model(drained behaviour)

The stiffness parameters of the Hardening Soil model are:

E50ref Secant stiffness in standard drained triaxial test [kN/m2]

Eoedref Tangent stiffness for primary oedometer loading [kN/m2]

Eurref Unloading / reloading stiffness (default Eurref = 3E50ref) [kN/m2]

νur Poisson's ratio for unloading-reloading (default ν = 0.2) [-]

Instead of entering the basic parameters for soil stiffness, alternative parameters can be entered. These
parameters are listed below:

CC Compression index [-]

Cs Swelling index or reloading index [-]

einit Initial void ratio [-]

PLAXIS 182 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

In addition, stress dependent stiffness can be defined through a power law with the following parameters:

m Power for stress-level dependency of stiffness [-]

pref Reference stress for stiffnesses (default pref = 100 kN/m2) [kN/m2]

The strength parameters of the Hardening Soil model coincide with those of the Mohr-Coulomb model:

c'ref Effective cohesion [kN/m2]

φ' Effective friction angle [°]

ψ Dilatancy angle [°]

In addition, a depth dependent cohesion can be defined:

c'inc As in the Mohr-Coulomb model (default cinc = 0) [kN/m3]

yref Reference level [m]

Dilatancy is generally dependent on the density of the soil: a loose soil shows less dilatancy than a dense soil. In a
shearing test it can be observed that if a dense material is sheared the effective dilatancy angle reduces with the
increase of volume strain, hence with reducing void ratio. On the other hand, in case of negative dilatancy, the
soil can only continue to reduce volume until the maximum density, hence the minimum void ratio, is reached.
PLAXIS 2D does not offer the possibility to take into account full void ratio dependent dilatancy, but it does have
the option to only calculate dilatancy if the void ratio is between certain limits while the dilatancy is considered
zero outside these limit values of the void ratio.

Dilatancy To be selected when dilatancy should only be considered within the void ratio [-]
cut-off range between emin and emax.

emin Minimum void ratio [-]

emax Maximum void ratio [-]

In some practical problems an area with tensile stresses may develop. This is allowed when the shear stress is
sufficiently small. However, the soil surface near a trench in clay sometimes shows tensile cracks. This indicates
that soil may also fail in tension instead of in shear. Such behaviour can be included in a PLAXIS 2D analysis by
selecting the Tension cut-off option. When selecting the Tension cut-off option the allowable tensile strength may
be entered. The default value of the tensile strength is zero.

Tension cut- To be selected when tension cut-off should be considered [-]


off

Tensile The allowable tensile strength [kN/m2]


strength

Finally, there are some Miscellaneous parameters to be defined that are not directly related to either stiffness or
strength. It is recommended to leave these values to their default values:

K0nc K0-value for normal consolidation (default K0nc = 1 - sin(φ)) [-]

PLAXIS 183 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Rf Failure ratio qf / qa (default Rf = 0.9) [-]

6.1.2.6 Hardening Soil model with small-strain stiffness (HSsmall)


Compared to the standard HS model, the Hardening Soil model with small-strain stiffness requires two
additional stiffness parameters as input: γ0.7 and G0ref. The Mechanical tabsheet for the Hardening Soil model
with small-strain stiffness is shown in Figure 127 (on page 185).

Note:
• Optional drainage types when Hardening Soil model with small-strain stiffness is selected are: Drained,
Undrained A, and Undrained B.
• In the case of Undrained A drainage type, the same parameters are used as for drained behaviour.
• In the case of Undrained B drainage type, φ = φu = 0, ψ = 0 and the undrained shear strength su is used instead
of the effective cohesion (c').

PLAXIS 184 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 127: Mechanical tabsheet for the Hardening Soil model with small-strain stiffness model (drained
behaviour)

All other parameters, including the alternative stiffness parameters, remain the same as in the standard
Hardening Soil model. In summary, the input stiffness parameters of the Hardening Soil model with small-strain
stiffness are listed below:
Parameters for stiffness:

PLAXIS 185 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

E50ref Secant stiffness in standard drained triaxial test [kN/m2]

Eoedref Tangent stiffness for primary oedometer loading [kN/m2]

Eurref Unloading / reloading stiffness (default Eurref = 3E50ref) [kN/m2]

m Power for stress-level dependency of stiffness [-]

Alternative parameters for stiffness:

CC Compression index [-]

Cs Swelling index or reloading index [-]

einit Initial void ratio [-]

Parameters for stress-dependent stiffness:

νur Poisson's ratio for unloading-reloading (default ν = 0.2) [-]

pref Reference stress for stiffnesses (default pref = 100 kN/m2) [kN/m2]

Parameters for small strain stiffness:

γ0.7 Shear strain at which Gs = 0.722 G0 [-]

G0ref Reference shear modulus at very small strains (ε < 10-6) [kN/m2]

General parameters for strength:

c'ref Effective cohesion [kN/m2]

φ' Effective friction angle [°]

ψ Dilatancy angle [°]

Parameters for depth-dependent strength:

c'inc As in Mohr-Coulomb model (default cinc = 0) [kN/m2/m]

yref Reference level [m]

Parameters to limit the dilatancy:

Dilatancy To be selected if a zero dilatancy angle should be used when the void ratio falls [-]
cut-off outside the limits defined by emin and emax.

emin Minimum void ratio [-]

emax Maximum void ratio [-]

PLAXIS 186 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Parameters for tensile strength:

Tension cut- To be selected when tension cut-off is considered [-]


off

Tensile The allowable tensile strength [kN/m2]


strength

Miscallaneous parameters:

K0nc K0-value for normal consolidation (default K0nc = 1 - sin(φ)) [-]

Rf Failure ratio qf / qa (default Rf = 0.9) [-]

In some practical problems an area with tensile stresses may develop. This is allowed when the shear stress is
sufficiently small. However, the soil surface near a trench in clay sometimes shows tensile cracks. This indicates
that soil may also fail in tension instead of in shear. Such behaviour can be included in a PLAXIS 2D analysis by
selecting the Tension cut-off option. When selecting the Tension cut-off option the allowable tensile strength may
be entered. The default value of the tensile strength is zero.
Hysteretic damping
The elastic modulus ratio is plotted as a function of the shear strain (γ) in a side pane when specifying the small-
strain stiffness parameters (Modulus reduction curve). The Hardening Soil model shows typical hysteretic
behaviour when subjected to cyclic shear loading. In dynamics calculations this leads to hysteretic damping. The
damping ratio is plotted as a function of the cyclic shear strain γC. Details are given in Brinkgreve, Kappert &
Bonnier (2007) (on page 552).

PLAXIS 187 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 128: Effect of small strain stiffness parameters on damping

Note:
Note that the Modulus reduction curve and the Damping curve are based on fully elastic behaviour. Plastic strains
as a result of hardening or local failure may lead to significant lower stiffness and higher damping.

6.1.2.7 Modified Cam-Clay model (MCC)


This is a critical state model that can be used to simulate the behaviour of normally consolidated soft soils. The
model assumes a logarithmic relationship between the void ratio and the mean effective stress. The Mechanical
tabsheet for the Modified Cam-Clay model is shown in Figure 129 (on page 189).

Note:
• Optional drainage types when Modified Cam-Clay model is selected are: Drained and Undrained A.
• In the case of Undrained A drainage type, the same parameters are used as for drained behaviour.

PLAXIS 188 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 129: Mechanical tabsheet for the Modified Cam-Clay model (drained behaviour)

Parameters for stiffness:

λ Cam-Clay compression index [-]

κ Cam-Clay swelling index [-]

νur Poisson's ratio for unloading/reloading [-]

Parameter for strength:

MCSL Tangent of the Critical State Line [-]

The Modified Cam-Clay model has an implicit relation between the ratio of horizontal and vertical stresses in
primary loading, K0NC, and the strength and stiffness parameters. Therefore in the Miscellaneous block the value
of K0NC resulting from the given stiffness and strength parameters is shown, but cannot be changed directly. If
independent control over the strength and stiffness parameters as well as K0NC is required, it is recommended to
use the Soft Soil model (SS) [ADV] (on page 191) instead of the Modified Cam-Clay model.

PLAXIS 189 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

6.1.2.8 NGI-ADP model


The NGI-ADP model may be used for capacity, deformation and soil-structure interaction analysis involving
undrained loading of clay-type materials. The Mechanical tabsheet for the NGI-ADP model is shown in Figure
130 (on page 190).

Note:
The only drainage type available when the NGI-ADP model is selected is Undrained C.

Figure 130: Mechanical tabsheet for the NGI-ADP model

Parameters for stiffness:

Gur/su,incA Ratio unloading/reloading shear modulus over the (plane strain) active shear [-]
strength

νu Undrained Poisson's ratio [-]

PLAXIS 190 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Parameters for strength

suA, ref Reference (plane strain) active shear strength [kN/m2]

su,incA Increase of (plane strain) active shear strength with depth [kN/m2/m]

yref Reference level for active, passive and DSS shear strength [m]

suC,TX/suA Ratio triaxial compressive shear strength over (plane strain) active shear strength [-]
(fixed at 0.99)

suP/suA Ratio of (plane strain) passive shear strength over (plane strain) active shear [-]
strength

suDSS/suA Ratio of direct simple shear strength over (plane strain) active shear strength [-]

γfC Shear strain at failure in triaxial compression (|γfC = 3/2 ε1C|) [-]

γfE Shear strain at failure in triaxial extension [-]

γfDSS Shear strain at failure in direct simple shear [-]

Miscellaneous parameters:

τ0/suA Ratio initial mobilized shear over (plane strain) active shear strength (default = [-]
0.7)

6.1.2.9 Soft Soil model (SS) [ADV]


The Mechanical tabsheet for the Soft Soil model is shown in Figure 131 (on page 192).

Note:
• Optional drainage types when Soft Soil model is selected are: Drained and Undrained A.
• For both Drained and Undrained A drainage type the same parameters are used.

PLAXIS 191 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 131: Mechanical tabsheet for the Soft Soil model (drained behaviour)

The parameters for stiffness are:

λ* Modified compression index [-]

κ* Modified swelling index [-]

νur Poisson's ratio for unloading-reloading (default ν = 0.15) [-]

Alternative parameters can be used to define stiffness:

CC Compression index [-]

Cs Swelling index or reloading index [-]

einit Initial void ratio [-]

The parameters for strength are:

PLAXIS 192 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

c'ref Effective cohesion [kN/m2]

φ' Effective friction angle [°]

ψ Dilatancy angle [°]

Tension cut- To be selected when tension cut-off is considered [-]


off

Tensile The allowable tensile strength [kN/m2]


strength

In some practical problems an area with tensile stresses may develop. This is allowed when the shear stress is
sufficiently small. However, the soil surface near a trench in clay sometimes shows tensile cracks. This indicates
that soil may also fail in tension instead of in shear. Such behaviour can be included in a PLAXIS 2Danalysis by
selecting the Tension cut-off option. When selecting the Tension cut-off option the allowable tensile strength may
be entered. The default value of the tensile strength is zero.
Miscellaneous parameters (use default settings):

K0nc K0-value for normal consolidation (default K0nc = 1 - sin(φ)) [-]

M K0nc- related parameter (no direct input possible) [-]

The parameter M is the inclination of the line passing through both the leftmost and topmost points of the ellipse
forming the hardening cap of the Soft Soil model. The value of M is derived from the value of K0nc and is given
here for comparison with the value MCSL of for instance the Modified Cam-Clay model (MCC) (on page 188).

6.1.2.10 Soft Soil Creep model (SSC) [ADV]


The Mechanical tabsheet for the Soft Soil Creep model is shown in Figure 132 (on page 194).

Note:
• Optional drainage types when Soft Soil Creep model is selected are: Drained and Undrained A.
• For both Drained and Undrained A drainage type, the same parameters are used.

PLAXIS 193 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 132: Mechanical tabsheet for the Soft Soil Creep model (drained behaviour)

The parameters for stiffness are:

λ* Modified compression index [-]

κ* Modified swelling index [-]

μ* Modified creep index [-]

νur Poisson's ratio for unloading-reloading (default ν = 0.15) [-]

Alternative parameters can be used to define stiffness:

CC Compression index [-]

PLAXIS 194 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Cs Swelling index or reloading index [-]

Cα Secondary compression index [-]

einit Initial void ratio [-]

The parameters for strength are:

c'ref Effective cohesion [kN/m2]

φ' Effective friction angle [°]

ψ Dilatancy angle [°]

Tension cut- To be selected when tension cut-off is considered [-]


off

Tensile The allowable tensile strength [kN/m2]


strength

In some practical problems an area with tensile stresses may develop. This is allowed when the shear stress is
sufficiently small. However, the soil surface near a trench in clay sometimes shows tensile cracks. This indicates
that soil may also fail in tension instead of in shear. Such behaviour can be included in a PLAXIS 2D analysis by
selecting the Tension cut-off option. When selecting the Tension cut-off option the allowable tensile strength may
be entered. The default value of the tensile strength is zero.
Miscellaneous parameters (use default settings):

K0nc K0-value for normal consolidation (default K0nc = 1 - sin(φ)) [-]

M K0nc- related parameter (no direct input possible) [-]

The parameter M is the inclination of the line passing through both the leftmost and topmost points of the ellipse
forming the hardening cap of the Soft Soil Creep model. The value of M is derived from the value of K0nc and is
given here for comparison with the value MCSL of for instance the Modified Cam-Clay model (MCC) (on page
188).

6.1.2.11 Sekiguchi-Ohta model (Inviscid) [ADV]


The Mechanical tabsheet for the Sekiguchi-Ohta model (Inviscid) is shown in Figure 133 (on page 196).

Note:
• Optional drainage types when Sekiguchi-Ohta model (Inviscid) is selected are: Drained and Undrained A.
• For both the Drained and Undrained A drainage type the same parameters are used.

PLAXIS 195 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 133: Mechanical tabsheet for the Sekiguchi-Ohta model (drained behaviour)

Parameters for stiffness:

λ* Modified compression index [-]

κ* Modified swelling index [-]

νur Poisson's ratio for unloading-reloading [-]

Alternative parameters can be used to define soil stiffnes:

CC Compression index [-]

Cs Swelling index or reloading index [-]

einit Initial void ratio [-]

The Sekiguchi-Ohta Inviscid model has a single strength parameter:

MCSL Tangent of critical state line [-]

PLAXIS 196 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Miscellaneous parameters:

K0nc Coefficient of lateral stress in normal consolidation [-]

6.1.2.12 Sekiguchi-Ohta model (Viscid) [ADV]


The Mechanical tabsheet for the Sekiguchi-Ohta model (Viscid) is shown in Figure 134 (on page 197).

Note:
• Optional drainage types when Sekiguchi-Ohta model (Viscid) is selected are: Drained and Undrained A.
• For both the Drained and Undrained A drainage type the same parameters are used.

Figure 134: Mechanical tabsheet for the Sekiguchi-Ohta model (Viscid) (drained behaviour)

The stiffness parameters of the Sekiguchi-Ohta model (Viscid) are:

PLAXIS 197 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

λ* Modified compression index [-]

κ* Modified swelling index [-]

νur Poisson's ratio for unloading-reloading [-]

Alternative parameters can be used to define soil stiffness:

CC Compression index [-]

Cs Swelling index or reloading index [-]

Cα* Secondary compression index [-]

einit Initial void ratio [-]

The model has a single parameter for soil strength:

MCSL Tangent of the critical state line [-]

Parameters for creep deformation:

α* Coefficient of secondary compression [-]

v̇ 0 Initial volumetric strain rate [day-1]

Miscellaneous parameters (using default is recommended):

K 0nc Coefficient of lateral stress in normal consolidation [-]

6.1.2.13 UDCAM-S model [ADV]


The UDCAM-S model can be used for capacity, deformation and soil-structure interaction analysis involving
undrained loading of clay-type materials for the design of offshore structures. The Mechanical tabsheet for the
UDCAM-S model is shown in Figure 135 (on page 199). The parameters are obtained from the Cyclic
accumulation and optimisation tool, which can be accessed by clicking on the corresponding button in the right-
hand panel.

Note:
The UDCAM-S model can only be used with the drainage type Undrained C.

PLAXIS 198 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 135: Mechanical tabsheet for the UDCAM-S model

Parameters for stiffness:

Gmax/τC Ratio unloading/reloading shear modulus over (plane strain) active shear strength [-]

νu Undrained Poisson's ratio [-]

Parameters for strength:

τrefC Degraded reference cyclic shear strength in triaxial compression [kN/m2]

τincC Increase of degraded cyclic shear strength in triaxial compression with depth [kN/m2/m]

yref Reference level for depth-dependent shear strength [m]

Ratio of degraded cyclic triaxial extension shear strength over degraded cyclic
τE/τC
triaxial compression shear strength

PLAXIS 199 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Ratio of degraded cyclic direct simple shear strength over degraded cyclic triaxial
τDSS/τC [-]
compression shear strength

τ0/τC Initial mobilisation [-]

γfC Shear strain at failure in triaxial compression [%]

γfE Shear strain at failure in triaxial extension [%]

γfDSS Shear strain at failure in direct simple shear [%]

Miscellaneous parameters:

τ0/τC Initial mobilisation [-]

6.1.2.14 Concrete model [ADV]


The Mechanical tabsheet for the Concrete model are shown in Figure 136 (on page 201).

Note:
Available drainage types when Concrete model is selected are: Drained, in case of semi-permeable walls, and
Non-porous, which is the general approach for concrete structural elements.

PLAXIS 200 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 136: Mechanical tabsheet for the Concrete model

PLAXIS 201 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Parameters for stiffness

E28 Young's modulus of cured concrete at thydr [kN/m2]

ν Poisson's ratio [-]

Parameters for strength:

Compression

fc,28 Uniaxial compressive strength of cured concrete at thydr [kN/m2]

fc0n Normalised failure strength [-]

fcfn Normalised initially mobilised strength [-]

fcun Normalised residual strength [-]

Gc,28 Compressive fracture energy of cured concrete at thydr [kN/m]

φmax Maximum friction angle [°]

ψ Dilatancy angle [°]

γfc Safety factor for compressive strength [-]

Tension

ft,28 Uniaxial tensile strength of cured concrete at thydr [kN/m2]

ftun Ratio of residual vs. peak tensile strength [-]

Gt,28 Tensile fracture energy of cured concrete at thydr [kN/m]

γft Safety factor for tensile strength [-]

The model has a single parameter for ductility:

aduct Increase of peak strain εcp with the increase of mean effective stress p [-]

Parameters for time-dependent behaviour:

thydr Time for full hydration [day]

Stiffness

Ratio of the Young's modulus after 1 day curing over the Young's modulus at full
E1 / E28
hydration

Strength

PLAXIS 202 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Strength
Strength development function for the first 24 hours of curing
function

Ratio of the compressive strength after 1 day curing over the compressive strength
fc,1 / fc,28 [-]
at full hydration

Ductility

εpcp,1h Uniaxial plastic failure strain at 1h (negative value) [-]

εpcp,8h Uniaxial plastic failure strain at 8h (negative value) [-]

εpcp,24h Uniaxial plastic failure strain at 24h (negative value) [-]

Shrinkage

Include
To be selected if shrinkage should be taken into account
shrinkage

ε∞shr Final shrinkage strain [-]

cr
t50 Time for 50% of shrinkage strains [day]

Creep

Include
To be selected if creep should be taken into account
creep

φ cr Ratio between creep and elastic strains [-]

cr
t50 Time for 50% of creep strains [day]

Note:
• When simulating shotcrete the tensile strength is essential for tunnel stability. Neglecting or considering low
values for tensile strength could result in unrealistic failure.
• The creep history is adjusted for the stress state at first activation of the concrete cluster, such that no creep
strains are produced by initial stresses. The state variables are taken over if the previous material was also
defined with the Concrete model, in which case creep will also continue. If a reset of state variables is desired,
a nil step with a different material (e.g. linear elastic) is required.

6.1.2.15 User-defined soil model (UDSM) [ADV] + [GSE]


The Mechanica tabsheet shows two drop-down menus; the top combo box lists all the DLLs that contain valid
User-defined Soil Model and the next combo box shows the models defined in the selected DLL. Each UD model
has its own set of model parameters, defined in the same DLL that contains the model definition.

PLAXIS 203 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

When an available model is chosen PLAXIS 2D will automatically read its parameter names and units from the
DLL and fill the parameter table below. For a detailed description of this facility, reference is made to the
Material Models Manual.

Note:
Available drainage types when user-defined soil model is selected are: Drained, Undrained A and Non-porous.

6.1.2.16 UBC3D-PLM model [ULT]


The Mechanical tabsheet for the UBC3D-PLM model is shown in Figure 137 (on page 205).

Note:
Available drainage types when UBC3D-PLM model is selected are: Drained and Undrained A.

PLAXIS 204 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 137: Mechanical tabsheet for the UBC3D-PLM model

The stiffness parameters of the UBC3D-PLM model are:

k B*e Elastic bulk modulus factor [-]

kG*e Elastic shear modulus factor [-]

kG* p Plastic shear modulus factor [-]

PLAXIS 205 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

me Power for stress dependent elastic bulk modulus [-]

ne Power for stress dependent elastic shear modulus [-]

np Power for stress dependent plastic shear modulus [-]

pref Reference pressure for stress dependent stiffness [kN/m2]

Note:
The implicit Poisson's ratio that is defined based on keB and keG is suitable for dynamics calculation, but it does
not generate a proper initial stress state if the initial stress condition is established by gravity loading procedure.
In such a case the user should define another material set for the stress initialization step with proper
characteristics.

Parameters for strength:

c Cohesion [kN/m2]

φcv Constant volume friction angle [°]

φp Peak friction angle [°]

σt Tensile strength [kN/m2]

The model requires a single parameter from field data in order to model the effects of densification:

(N1)60 Corrected SPT blow count [-]

Advanced parameters:

fdens Densification factor [-]

fEpost Post-liquefaction stiffness factor [-]

Rf Failure ratio [-]

6.1.2.17 Parameters for Excess pore pressure calculation


The excess pore pressure calculation parameters available in the Mechanical tabsheet can be used to model the
generation of excess pore pressures in a calculation that involves the Undrained behaviour of soils or a coupled
analysis. For the calculation of excess pore pressures in an effective stress calculation, the calculation will add a
bulk modulus of water to the stiffness matrix. This bulk modulus can be obtained in 2 different ways (Figure 138
(on page 207)), namely using the options ν-undrained definition and Biot effective stress concept. When the first
option is chosen the user should decide whether to enter the νu,equivalent either directly or through the Skempton
B value. Note that we refer here to the equivalent undrained Poisson's ratio as it is not directly used as
parameter to calculate excess pore pressures, but merely to determine the bulk modulus of water that is then
used to calculate excess pore pressures.

PLAXIS 206 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

By default, the solid material of the soil (i.c the grains) is supposed to be incompressible (αBiot = 1) and the
equivalent undrained Poisson's ratio is taken as 0.495. This is the default behaviour. Alternatively, one can select
to define the equivalent undrained Poisson's ratio through the Skempton's B-parameter and Kw,ref/n is calculated
accordingly. In the most advanced option, Biot effective stress concept, the compressibility of the solid material is
considered and one can select αBiot together with Kw after which νu,equivalent, Skempton B and Kw,ref/n are
calculated accordingly.

αBiot Biot alpha pore pressure coefficient [-]

νu,equivalent Equivalent undrained Poisson's ratio [-]

Skempton-B A parameter that determines which portion of a change in mean stress is carried [-]
by the pore water

Kw,ref / n The corresponding reference bulk stiffness of the pore fluid [kN/m2]

Kw Bulk modulus of water [-]

More detailed information is available in the Material Models Manual.

Figure 138: Parameters for the Excess pore pressure calculation

6.1.3 Hoek-Brown pre-processing tool


The Hoek-Brown model is the most used failure criterion for rock masses, nevertheless there are some
uncertainties regarding the input parameters that require a consolidated experience (Hoek, Carranza-Torres &
Corkum (2002); Hoek (2007)). For this reason, PLAXIS 2D implements in the Mechanical tabsheet of the Hoek-
Brown model a pre-processing tool to guide the user in the determination of the rock mass strength and stiffness
parameters. PLAXIS 2D sign convention is adopted, i.e. compressive stresses are considered to be negative. The
pane contains the following tabsheets:

PLAXIS 207 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

• Analysis: shows the Hoek-Brown failure envelope in the plane of principal effective stresses σ'3 - σ'1, in order
to visualise the effects of changing of rock mass parameters on the failure envelope. For more information
about the Hoek-Brown modeland its formulation, refer Material Models Manual.
• The second tabsheet is specific to determine the selected parameter.

6.1.3.1 Uni-axial compressive strength of intact rock |σci|


The uni-axial compressive strength of intact rock |σci| can be checked based on the estimation methods generally
executed in the field. Alternatively, a range can be selected and PLAXIS 2D automatically set the parameter to the
middle value of the range. In case the upper bound limit of 250 MPa is selected, the parameter is set to 250 MPa.
Table 4 (on page 208) reports the range of compressive strength for each category with a qualitative description
of the strength behaviour to field test.

Table 4: |σci| tabsheet values

Field estimate Examples Strength [MPa]

Only chipping is possible with a geological Chert, diabase, fresh basalt, gneiss,
|σci| ≥ 250
hammer granite, quartzite

Amphibolite, basalt, gabbro, gneiss,


Fracturing requires many blows of a
granodiorite, limestone, marble, 100 ≤ |σci| ≤ 250
geological hammer
rhyolite, sandstone, tuff

Fracturing requires more than one blow of a Limestone, marble, phyllite,


50 ≤ |σci| ≤ 100
geological hammer sandstone, schist, shale

Fracturing is possible with a single blow


Claystone, coal, concrete, schist, shale,
from a geological hammer, but cannot be 25 ≤ |σci| ≤ 50
siltstone
scraped or peeled with a pocket knife

Firm blow with the point of a geological


hammer leaves shallow indentation; peeling Chalk, potash, rocksalt 5 ≤ |σci| ≤ 25
with a pocket knife is possible, but difficult

Firm blow with the point of geological


hammer leads to crumbling; peeling with a Highly weathered or altered rock 1 ≤ |σci| ≤ 5
pocket knife is possible

Thumbnail leaves indentation Stiff fault gouge 0.25 ≤ |σci| ≤ 1

6.1.3.2 Intact rock parametermi


The intact rock parameter mi can be estimated based on the rock type.

PLAXIS 208 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Table 5: Values of the constant mi for intact rock (Marinos & Hoek (2001), Wyllie & Mah (2004))

Name Rock type Texture mi mi \pm

Agglomerate Igneous Coarse 19 3

Amphibolites Metamorhpic Medium 26 6

Andesite Igneous Medium 25 5

Anhydrite Sedimentary Fine 12 2

Basalt Igneous Fine 25 5

Breccia Igneous Medium 19 5

Breccia Sedimentary Coarse 19 5

Chalk Sedimentary Veryfine 7 2

Claystones Sedimentary Veryfine 4 2

Conglomerates Sedimentary Coarse 21 3

Crystallinelimestone Sedimentary Coarse 12 3

Dacite Igneous Fine 25 3

Diabase Igneous Fine 15 5

Diorite Igneous Medium 25 5

Dolerite Igneous Medium 16 5

Dolomites Sedimentary Veryfine 9 3

Gabbro Igneous Coarse 27 3

Gneiss Metamorhpic Fine 28 5

Granite Igneous Coarse 32 3

Granodiorite Igneous Coarse,Medium 29 3

Greywackes Sedimentary Fine 18 3

Gypsum Sedimentary Medium 8 2

Hornfels Metamorhpic Medium 19 4

Marble Metamorhpic Coarse 9 3

Marls Sedimentary Veryfine 7 2

PLAXIS 209 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Name Rock type Texture mi mi \pm

Metasandstone Metamorhpic Medium 19 3

Mictriclimestones Sedimentary Fine 9 2

Migmatite Metamorhpic Coarse 29 3

Norite Igneous Coarse,Medium 20 5

Obsidian Igneous Veryfine 19 3

Peridotite Igneous Veryfine 25 5

Phyllites Metamorhpic Fine 7 3

Porphyries Igneous Coarse,Medium 20 5

Quartzites Metamorhpic Fine 20 3

Rhyolite Igneous Medium 25 5

Sandstones Sedimentary Medium 17 4

Schists Metamorhpic Medium 12 3

Shales Sedimentary Veryfine 6 2

Siltstones Sedimentary Fine 7 2

Slates Metamorhpic Veryfine 7 4

Spariticlimestones Sedimentary Medium 10 2

Tuff Igneous Fine 13 5

6.1.3.3 Geological Strength Index GSI


The GSI tabsheet allows to choose between two rock types: General Figure 139 (on page 211) and Flysch Figure
140 (on page 212). An intact rock is equivalent to GSI = 100, whereas a soil structure is in proximity to GSI = 0.
Based on the selected rock type, the GSI can be chosen taking the structure and the surface conditions of the rock
mass into account. An hint box at the bottom of the chart displays the rock characteristics for the selected GSI.

Note:
The GSI charts allows only integer values. If a higher accuracy is required, a decimal value can be manually put in
the Input tabsheet of the material set.

PLAXIS 210 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 139: GSI chart for General rock type

PLAXIS 211 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 140: GSI chart for Flysch rock type

6.1.3.4 Disturbance factor (D)


The disturbance factor D depends on the amount of disturbance of the rock as a result of mechanical processes
in open excavations, tunnels or mines, such as blasting, tunnel boring, machine driven or manual excavation. No
disturbance is equivalent to D = 0, whereas severe disturbance is equivalent to D = 1. The D tabsheet allows a
direct estimation of the disturbance factor for tunnel Table 6 (on page 213) and slopes Table 7 (on page 213).
In blast damaged rock, there are two main issues when applying the disturbance factor D:
• Choose a suitable value: a large number of factors can influence the degree of disturbance (quality of the rock
mass and the excavation/blasting, loading sequence, lateral confinement produced by different radii of
curvature of slopes as compared with their height, the level of strain in the failure zone, etc.). For this reason,
the disturbance factor D is inferred on the basis of engineering experience and research Table 6 (on page
213)Table 7 (on page 213).
• Define the extent of the damaged zone: the disturbance factor D should only be applied to the actual zone of
damaged rock in order to avoid underestimation of the strength and stability of the overall rock mass greatly.
The thickness T of the blast damaged zone depends upon the design of the blast (Hoek & Karzulovic, 2000).
In the case of bench blasting in open pit mines and civil engineering slopes, Table 7 (on page 213) contains
suggested values, where H is the height of the slope. In the case of tunnels excavated by drill and blast
methods, for very high quality controlled blasting the damage to the tunnel wall is negligible due to a well-
designed blasting pattern, detonation sequence and accurate drilling control. In contrast, the lack of a good
blast design and absence of any control on the drilling can result in significant damage.

PLAXIS 212 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Table 6: Distrubance factor D, guidelines for tunnels

Description D

Tunnel excavation by TBM or blasting of excellent quality 0.0

Tunnel excavation by hand or using a mechanical process rather than blasting, in poor quality
rock. There are no squeezing problems leading to floor heave, or these are mitigated using a 0.0
temporary invert

Tunnel excavation by hand or using a mechanical process rather than blasting, in poor quality
0.5
rock. There are unmitigated squeezing problems leading to floor heave

Tunnel excavation using blasting of very poor quality, leading to severe local damage 0.8

Table 7: Disturbance factor D, guidelines for slopes

Description D

Slope created using controlled, small scale blasting of good quality 0.7

Slope created using small scale blasting of poor quality 1.0

Slope in very large open pit mine, created using mechanical excavation in softer rocks 0.7

Slope in very large open pit mine, created using heavy production blasting 1.0

Table 8: Damaged thickness T, guidelines for bench blasting in open pit mines and civil engineering slopes

Description T

Large production blast, confined and with little or no control 2.0 to 2.5 H

Production blast with no control but blasting to a free face 1.0 to 1.5 H

Production blast, confined but with some control, e.g. one or more buffer rows 1.0 to 1.2 H

Production blast with some control, e.g. one or more buffer rows, and blasting to a free face 0.5 to 1.0 H

Carefully controlled production blast with a free face 0.3 to 0.5 H

6.1.4 Cyclic accumulation and optimisation tool [ADV]


The Cyclic accumulation and optimisation tool is used to determine the UDCAM-S model parameters. The model
is especially suited for the foundation design of offshore structures subjected to a design storm (i.e. a
combination of a wave, wind loading, and currant). The stiffness and strength of saturated soils under cyclic
loading are different from the static case: they may increase due to strain rate effects or reduce due to
degradation processes, as pore pressure build up and destructuration. The soil beneath structures subjected to

PLAXIS 213 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

cyclic loading experiences different behaviour following different stress paths. For a general stress path, the
behaviour may be interpolated between triaxial compression (TxC), triaxial extension (TxE) and direct simple
shear (DSS) conditions.
The Cyclic accumulation and optimisation tool offers the possibility to check the effect of a specified input load
history for a specific soil type and, through an accumulation and interpolation procedure, it provides stress-
strain curves for the different characteristic stress paths. The UDCAM-S model parameters are optimised in
order to match the stress-strain curves.
The tool is suitable only for clay and low permeable silts in undrained condition since the effect of drainage is
not taken into account. The accumulation procedure is based on the cyclic shear strains.

Note:
In addition to the cyclic shear strain accumulation procedure, the pore pressure accumulation and average shear
strain accumulation procedures can be found in literature. The reason for using different procedures is that the
development of one of these parameters may be more prominent than the others, depending on soil type, static
and cyclic shear stresses, etc. (Andersen, 1991 (on page 552)). The cyclic shear strain accumulation procedure
has been extensively used for clays, while the pore pressure accumulation procedure has been used for sands
since drainage is likely to occur during the complete load history (even though not within each cycle) in this type
of soils. In principle, the pore pressure accumulation procedure could also be used for clays. In practice,
however, accurate laboratory measurement of pore pressure is harder to perform in clays than in sand. Since
drainage is not likely to occur during the cyclic load history in clays, it is preferable to use the cyclic strain
accumulation procedure (Andersen, 2004 (on page 552)).

6.1.4.1 Starting the tool


To start the tool, the Cyclic accumulation and optimisation tool is clicked from the side panel in the Mechanical
tabsheet of the UDCAM-S model . The window consists of two blue tabs and one orange tab, see Figure 141 (on
page 215) for an overview.

PLAXIS 214 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 141: Cyclic accumulation and optimisation tool

The first blue tabsheet, Cyclic accumulation, gives information regarding the input load history and the soil type.
The second blue tabsheet, Stress-strain curves, gives information about the stress-strain curves for different
stress paths. The orange tabsheet Parameter optimisation, allows the user to optimise the UDCAM-S model
parameters.
Above the tabsheets there are Load and Save buttons (see Figure 142 (on page 215)) to load or save the app
state. Saving and loading the app state is useful when running the same or similar calculations again, after the
project has been closed and reopened. Also, similar calculations in different projects can be done more easily by
loading a previously saved state.

Figure 142: Load and save app state

The state is saved in the JSON file format with the .json extension. All the settings and data on the tabsheet, like
table values and graph data are saved and can be loaded back into the tool.

PLAXIS 215 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

6.1.4.2 Cyclic accumulation


In the design process of offshore structures, the design load is often a storm which can be transformed or
idealised into packages or parcels. Each parcel corresponds to a number of cycles at a load constant amplitude.
The maximum cyclic load and the number of cycles at each load level are determined from a time record of loads.
The counting procedure to obtain the number of cycles can be done following different methods (e.g. the
rainflow-counting algorithm (Matsuishi & Endo (1968) (on page 553))) or the recently developed method, for
more information see Noren - Cosgriff, Josatd & Madshus (2015) (on page 553)).
The effect of the design storm on the soil deposit can be described through the equivalent number of cycles Neq
associated to the last applied normalised load amplitude F/Fmax and the soil cyclic contour diagram. The
equivalent number of cycles Neq is calculated at failure for capacity calculations. If it is found to be conservative,
the same value may also be used in stiffness calculations.

Note:
It is assumed here that Neq is representative for all the elements in the selected soil polygon. Even though this
assumption may underestimate the effect of the stress redistribution and progressive failure, it has been shown
that the results are in good agreement with the ones from the model tests (Andersen, Puech & Jardine, 2013 (on
page 552)) and can certainly be used for stiff shallow foundations like offshore structures, while for monopiles
and piles this is generally not correct.

The cyclic contours (Figure 143 (on page 217)) are a set of curves where each of them represents the locus of
points characterised by the same cyclic shear strain. They can be visualized in a chart with the number of cycles
on the x-axis (in a logarithmic scale) and τcyc/su on the y-axis. Each point of the contour is then characterised by
a certain number of cycles and a normalised shear stress that induce a defined cyclic shear deformation in the
soil. The contours are defined for DSS tests with the average stress τa equals to 0 (for Drammen clay).

PLAXIS 216 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 143: Contour diagram for DSS test for Drammen Clay with OCR = 4

Four lines limit the contours:


• Each contour line starts with a point at N equal to 1 and an ending point at N equal to the number of cycles
above which all the contours are assumed to become horizontal (i.e. τcyc is constant). This leads to two
vertical boundaries, N = 1 and N = Nmax.
• The first contour is always defined at zero stress level, i.e. the contour line at the base where τcyc/su is equal
to zero and γcyc = 0%.
• The last contour represents the locus of points corresponding to γ equal to the cyclic shear strain at failure
(typically 15%).
Contour diagram data
The cyclic behaviour of clays depends on OCR, as well as the other index properties such as Plasticity Index, clay
content, water content, etc. NGI has extensively studied the cyclic behaviour of Drammen clays, realizing contour
diagrams from DSS tests. These contour data for Drammen clay with OCR equal to 1, 2, 4, 10, 20 and 40 are
available in the tool and can be selected from the drop-down menu.
In the case of other clays, it is possible to upload custom contours. In general, the contours can be determined by
different tests at different average shear stress levels, τa. For Drammen clay, Neq was found to be relatively
independent of τa and the type of test (Andersen, Kleven & Heien, 1988 (on page 552)) and the shear strain at
failure was taken as 15%.

Note: When uploading custom contours, a specific data format has to be used. For more information, see Cyclic
accumulation Tool - Contour File Formats [ADV] (on page 558).

PLAXIS 217 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

By moving the mouse over any line in the contour chart, a hint box appears showing the value of γcyc that
describes that contour and the corresponding Ncycles and τcyc/su values.

Note:
• The shear strain at a given degree of shear strength mobilisation, τ / su, increases with increasing OCR,
meaning that the normalised secant modulus, G/su, decreases with increasing OCR for a given degree of
strength mobilisation. However, the absolute value of the stiffness will increase with increasing OCR for given
consolidation stress and strength mobilisation but significantly less than the increase in shear strength
(Andersen, 2015 (on page 552)).
• Even though in most cases the strength of the soil under cyclic loading may be reduced compared to the static
value, because of the degradation process, the cyclic shear strength can also be higher than the static shear
strength (i.e. τcy / su > 1) for a low number of cycles in some cases. The reason for this is that the clay strength
is rate dependent. Since the cyclic tests are typically run with a load period of 10 seconds and the monotonic
tests are brought to failure in about 2 hours, the cyclic strength may thus be higher than the static shear
strength (Andersen, 2007 (on page 552)).

Load ratio (F/Fmax)


The idealised load history, described by a series of parcels, is specified in a table that consists of four columns.
The first one indicates the number of the load parcel, the second one represents the load ratio value F / Fmax for
each parcel, the third one shows the corresponding number of cycles and the last one is the output of the
calculation performed in this tabsheet (i.e. the stress ratio). The load ratio F/Fmax refer to pure cyclic loading
condition (horizontal force, vertical force or bending moment). The user should define which load direction is
more crucial for the model and calculate consequently F/Fmax. For capacity analysis, when F is equal to Fmax (i.e.
the load ratio is equal to 1), failure has been reached and the shear strain is equal to its failure value γf. The
normalized shear strength τ/su at γf corresponds to its maximum value. A linear relationship between the load
and the shear stress is assumed, so that:

( )
τ
su max
= χF
F
max
Eq. [15]

where χ is the scaling factor automatically calculated during the strain accumulation procedure. At the end of the
accumulation procedure, the maximum stress ratio τ/su is displayed in the corresponding column in the Figure
144 (on page 219) .
The buttons in the toolbar can be used to modify the Figure 144 (on page 219).

PLAXIS 218 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 144: Input table for load parcels of the cyclic accumulation and optimisation tool


Click the Add row button in the toolbar to add a new row in the table.

Click the Insert button to insert a new row before the selected row in the table.

Click the Delete button to delete the selected row in the table.
The values can be defined by clicking the cell in the table and by typing the value. A maximum of 100 load
parcels can be entered. The load ratio can be defined as any positive value larger than zero; its maximum value is
1. The number of cycles is a positive integer larger than zero.
As soon as a parcel is specified in the table, the graph "Load ratio vs N cycles'' at the bottom of the same tab will
be updated. There is an option to choose for a logarithmic representation on the y-axis. See an example of where
a series of parcels were defined in Figure 145 (on page 220).

If the data for the load parcels exists in a file, it can be loaded by clicking the Open file. Any delimited
tabular text file or an excel spreadsheet file can be loaded. When the selected file is opened, a window pops
up where the user can choose the columns to extract the data from, see Figure 144 (on page 219). Also, the
field separator (delimiter) and start row can be defined. A preview of what is going to be imported is shown.

A load history, defined in the table, can be copied using the Copy button in the toolbar.

Copied data from other applications (using Ctrl+c) can be imported by using the Paste button.

PLAXIS 219 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 145: Automatic update of Load ratio graph when specifying parcels

Note:
If a load parcel is characterized by a number of cycles larger than the maximum value represented in the
contours diagram, it is advised to decompose it into two or more load parcels with slightly different load ratios
and lower number of cycles such that their sum is equal to the initial total Ncycles.

Figure 146: Import load parcels settings dialog

PLAXIS 220 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

To start the strain accumulation procedure, click the Calculate button. Prior to the calculation, the tool performs
some consistency checks on the input data to ensure that both the load parcel table and the contour diagram are
defined.
At the end of the calculation, the box of the equivalent number of cycles Neq is updated with the value resulting
from the strain accumulation procedure. The last accumulation up to failure is now visible in the contour
diagrams together with the locus of the end points for each accumulation (i.e. for all scaling factors used in the
process).
Cyclic accumulation tab possible errors
The errors that may occur in the Cyclic accumulation tab are listed in Calculation warning and errors in PLAXIS
(on page 625).

6.1.4.3 Stress-strain curves


The Stress-strain curves tabsheet allows to determine the stress-strain curves for different stress paths. The
curves are determined based on contour diagrams for different laboratory test types at a given equivalent
number of cycles. It is possible to choose among several Drammen clay contours (for OCR = 1, 2, 4, 10, 20 and
40) or to upload custom diagrams.

Note: When uploading custom contour diagrams, a specific data format has to be used. Please see Cyclic
accumulation (on page 558) for more information

The Drammen clay contours have been built as result of a large DSS and triaxial cycling testing programme
(Andersen et al, 1988 (on page 552)). Each point in the diagram is described by the average and cyclic shear
stresses, τa and τcyc, under which the tests were run, normalised with respect to the static undrained shear
strength in triaxial compression suC, together with the number of cycles and the corresponding value of the
average and cyclic shear strain at failure. For Drammen clay, failure is considered to be reached when either γa
or γcyc is equal to 15%. The total maximum shear stress τf, cyc that can be mobilized is given by the sum of the
average and the cyclic shear stress. In each diagram, the cyclic shear strength is determined for a case with a
constant cyclic shear stress during the cyclic load history. In reality, during a storm, the cyclic shear stress varies
from one cycle to the next one. The value of Neq is therefore essential to use these diagrams. By grouping the test
results for selected number of cycles, it has been established contour diagrams for N = 1, 10, 100, 1000. The
Stress-strain curves tabsheet automatically interpolates the contours for the desired Neq, representative of the
real cyclic load history.

PLAXIS 221 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 147: Stress-strain curves tabsheet, with contour diagrams for Drammen clay with OCR = 4.

By default, the equivalent number of cycles is copied from the Cyclic accumulation tabsheet. The contour
diagrams for DSS and triaxial tests are selected to be consistent with the chosen ones to perform the strain
accumulation procedure. If a custom contour has been chosen, it is necessary to upload a custom contour for the
strain interpolation procedure (describe format of the file). This can be done by setting Neq determination>
Manual. Then the option Contour diagrams will enable to Select contour diagrams by selecting Custom contours
from the drop-down list.

Note:
For an advanced use of the tool, it is possible to select the Manual option from the Neq determination drop-down
menu. This allows to manually enter a value for Neq or to modify the existing one, as well as selecting another
contour diagram from the list in the corresponding drop-down menu.

If the Soil behaviour is Anisotropic, i.e. the stress-strain curves are determined both for DSS and for triaxial tests,
it is generally expected that the anisotropy ratios τDSS/τC and τE/τC are different from 1. On the contrary, when
choosing the Isotropic option in the corresponding drop-down menu, the strain interpolation is performed for
the DSS test only and the anisotropy ratios are all equal to 1.
Two different scaling factors can be used to scale the y-axis of the DSS and triaxial diagrams, respectively.
The stress-strain curves are determined by specifying the ratio of the cyclic shear stress to the average shear
stress for each test condition (DSS, TXC, TXE) according to the desired failure mode and loading condition. When
a stress ratio different from zero is defined, the path line is automatically updated in the corresponding chart.
Note that the starting point of the path line is always on the x-axis (τcyc/suC and γcyc equal to zero) but it does not
always coincide with the origin of the diagram: it is represented by the intersection point between the x-axis and
the contour line that corresponds to γa = 0%. For the DSS test, when the average strain is equal to zero also

PLAXIS 222 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

τa/suC is equal to zero, while for the triaxial tests, the intersection generally occurs at τa/suC different from zero.
In fact, it corresponds to the initial in situ mobilization τ0/suC.

Note:
The contour diagram shows that the maximum cyclic shear strength is mobilized when large cyclic shear strains
occur and when the average shear stress is small. On the contrary, the maximum average shear strength is
mobilized when the average shear stress approaches the static shear strength and large average shear strains
occur. For average shear stresses between zero and the static shear strength, the failure mode is a combination
of average and cyclic shear strains.

The stress ratios should be defined such that (Δτcyc / Δτa)DSS is representative of the failure mode or, in case of
difference between the application direction of the cyclic and average load (e.g. one acts mostly in the horizontal
direction and the other one mainly in the vertical), it is advised to calculate τcyc/τa as the ratio between stresses
representative for the boundary value problem. The stress ratios for the triaxial stress may be chosen in order to
have strain compatibility with the DSS test, i.e. similar average and cyclic shear strains must be reached. As
shown in the chart, since each stress ratio represents the inclination of the corresponding stress path line, a
negative value must be entered for (Δτcyc/Δτa)TXE and a positive one for the others.
For establishing stress-strain curves, it is possible to choose among three load types:
• Cyclic load: the calculation follows the actual ratio between τcyc and τa (inclined path), but only the
relationship between τcyc and γcyc is considered.
• Average load: the calculation follows the actual ratio between τcyc and τa (inclined path), but only the
relationship between τa and γa is considered.
• Total load: the calculation considers the total strength, (τcyc + τa) and the total shear strain, (γcyc + γa).
To start generating stress-strain curves, click the Calculate button. Prior to the calculation, the tool performs
some consistency checks on the input data to ensure that both a valid equivalent number of cycles (larger than
0) and a contour diagram have been selected.

PLAXIS 223 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 148: Stress-strain curves tabsheet displaying results after calculation

The Target curves chart shows the stress-strain curves for all tests (e.g. only DSS if the isotropic behaviour has
been chosen) and for the selected load type. On the x-axis the shear strain γ is plotted in percentage, while the y-
axis corresponds to the normalised shear strength τ / suC. The curves are used as input to the third and last
tabsheet Parameter optimisation.

6.1.4.4 Parameter optimisation


In the Parameter optimisation tabsheet (Figure 149 (on page 225)) the UDCAM-S model optimised parameters
can be determined by simulating undrained DSS, TxC, and TxE tests on one material point. The optimisation
procedure is based on the Particle Swarm Algorithm that allows to find the parameter values resulting in a
stress-strain curve that matches the target one. The laboratory test conditions are determined based on the
Static properties table.

PLAXIS 224 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 149: Parameter optimisation tabsheet

The parameters with their standard units are listed below.


Static properties:

su,refC Reference undrained triaxial compression TxC shear strength [kN/m2]

yref Reference depth [m]

su,incC Increase of TxC shear strength with depth [kN/m2/m]

τ0/suC Initial mobilization [-]

suC / σ'yy Ratio of the undrained compression shear strength over the current vertical [-]
effective stress

K0 Lateral earth pressure coefficient at rest [-]

The initial mobilization is determined based on the contour diagrams in the Stress-strain curves tabsheet: it
corresponds to the value of τa / suC at the intersection point between the x-axis (γcyc = 0%) and the contour line
for γa equal to 0%.
The K0 determination option is set by default to automatic. This means that the value of K0 used to determine the
initial stress condition of the test is determined based on the following equation:

(
K0 = 1 − 2
τ0 suC

suC σ ′ yy
) Eq. [16]

PLAXIS 225 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Note:
If in the Stress-strain curves tabsheet the user has selected the Cyclic load type, the initial mobilisation and the
undrained shear strength normalised by the vertical effective stress cannot be modified and are equal to 0 and 1,
respectively.

In the Parameter ranges table it is required to define the minimum and maximum value that should be
considered for each parameter during the optimisation procedure. The ranges can be estimated based on the
target charts.
Double-clicking on one of the graphs opens the selected chart in a bigger window (Figure 150 (on page 226)).
This window shows the selected diagram, the table of the data points that are used to plot it as well as the
tangent and the secant values of the plot. Note that the point to be taken into consideration for the calculation of
the tangent and the secant values can be determined by clicking on the plot. The secant and the tangent are
useful for the back-calculation of stiffness parameters from stress-strain diagrams. The corresponding secant
and tangent values are indicated below the table. The graph or the data can be copied to the clipboard by
selecting the corresponding option in the drop-down menu displayed when the Copy button is clicked. The
diagram can be zoomed in or out using the mouse by first clicking and holding the left mouse button in the
diagram area and then moving the mouse to a second location and releasing the mouse button. Moving the
mouse from the left upper corner to the right lower corner zooms the diagram to the selected area, whereas
moving the mouse from the right lower corner to the left upper corner resets the view. The zoom action can also
be undone using the Zoom out option on the toolbar. The wheel button of the mouse can be used for panning:
click and hold the mouse wheel down and move the diagram to the desired position.

Figure 150: Chart displaying target, optimised, secant and tangent curves, and corresponding table of values.

Parameter ranges:

PLAXIS 226 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Gmax/τC Ratio of the initial shear modulus to the cyclic compression shear strength [-]

γfC Shear strain at failure in triaxial compression [%]

γfE Shear strain at failure in triaxial extension [%]

γfDSS Shear strain at failure in direct simple shear [%]

τC/suC Ratio of the cyclic compression shear strength over the undrained static [-]
compression shear strength

τE/suC Ratio of the cyclic extension shear strength over the undrained static compression [-]
shear strength

The initial shear modulus defines the tangent value of the stress strain curves at the initial shear stress
(Anderson, 2004 (on page 552)). For high loading levels, the area where G is equal to Gmax is very limited and
therefore not governing. For most problems, Gmax/suC has to be chosen to fit the stress-strain curve in the actual
strain range of interest. In the case of isotropic behaviour, γfC and γfE are set equal to the optimised γfDSS, while
the ratios τDSS/τC and τE / τC are equal to 1.
To run the optimisation procedure, click the Calculate button.
The results are shown both in the Parameter ranges table and in the charts. The optimum values of the
parameters used to obtain the best fit to the test data are shown in the Optimal value column of the table. If the
optimum value is equal to the minimum or maximum value, it might be that the best value lies outside the
specified range. The last column of the table shows the sensitivity of each parameter. A sensitivity of 100%
means that the parameter has a high influence on the simulated test results, whereas a low sensitivity values
means that the parameter has a low influence on the simulated test results.
The optimised curves are shown in the corresponding test graph. As explained above, each chart can be opened
in a larger window by double-clicking it. Also for the optimised curve the table with the data points and the
possibility to draw the secant and tangent lines are available.
The optimal values are used to determine the optimised parameters used in the UDCAM-S model and shown in
the Optimised parameters table (Figure 149 (on page 225)).

PLAXIS 227 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 151: Optimised Parameters for Cyclic accumulation and optimisation tool

The parameters with their standard units are listed below.


Derived parameters:

Gmax/τC Ratio of the initial shear modulus to the cyclic compression shear strength [-]

γfC Shear strain at failure in triaxial compression [%]

γfE Shear strain at failure in triaxial extension [%]

γfDSS Shear strain at failure in direct simple shear [%]

τrefC Cyclic compression shear strength [kN/m2]

yref Reference depth [m]

τincC Increase of the cyclic shear strength with depth [kN/m2]

τE/τC Ratio of the cyclic extension shear strength over the cyclic compression shear [-]
strength

τ0/τC Cyclic mobilization [-]

τDSS/τC Ratio of the cyclic DSS shear strength over the cyclic compression shear strength [-]

For more information about the parameters of the UDCAM-S model , reference is made to Material Models
Manual - UDCAM-S.
To copy the Optimised parameters table to the material database, click the Copy parameters.

PLAXIS 228 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Possible errors of Parameter optimisation tab


For more information on the errors that may occur in the Parameter optimisation tab, refer to Table 27 (on page
633).

6.1.5 Groundwater tabsheet


Flow parameters are required when dealing with problems that involve flow of pore water in saturated or
unsaturated soils, i.e. when using Groundwater flow, Consolidation or Fully coupled flow-deformation types of
calculation. When considering steady-state groundwater flow or consolidation of fully saturated soil layers, only
the soil's (saturated) permeability is a relevant parameter. However, when considering unconfined flow,
seepage, transient (time-dependent) flow or fully coupled flow-deformation analysis, partially saturated soil
behaviour becomes an issue and needs to be described in more detail. This requires, amongst other things, the
selection of a so-called soil-water retention curve relating the suction (positive pore water stress) in the
unsaturated zone to the degree of saturation.
PLAXIS 2D incorporates functions to describe the flow behaviour in the unsaturated zone, among which the
famous Mualem-Van Genuchten functions. In order to enable an easy selection of the unsaturated flow
parameters in these functions, predefined data sets are available for common types of soil. These data sets can
be selected based on standardised soil classification systems.

Note:
Although the predefined data sets have been created for the convenience of the user, the users remain at all
times responsible for the model parameters that they use. Note that these predefined data sets have limited
accuracy.

6.1.5.1 Hydraulic data sets and models


The program provides different data sets and models to model the flow in the saturated zone in soil. The data
sets available in the program are:
1. Standard:
This option allows for a simplified selection of the most common soil types (Coarse, Medium, Medium fine,
Fine and Very fine non-organic materials and Organic material) and is based on the Hypres topsoil
classification series (Wosten, Lilly, Nemmes & Bas, 1999 (on page 554)). The only model available for this
data set is Van Genuchten (see Material Models Manual - Hydraulic Models).
When one of the soil type options is selected, the particle fractions are automatically defined and the soil type
is indicated in the soil texture triangle (Figure 152 (on page 230)). The particle fractions can also be defined
by clicking on the corresponding location in the soil texture triangle or by directly typing the values.

PLAXIS 229 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 152: Groundwater parameters for Standard data set (Wösten, Lilly, Nemmes & Bas, 1999)

2. Hypres:
The Hypres series is an international soil classification system (Wosten, Lilly, Nemmes & Bas, 1999 (on page
554)). The hydraulic models available for Hypres data set are the Van Genuchten model and the Approximate
Van Genuchten (see Material Models Manual - Hydraulic Models).
A distinction can be made between Topsoil and Subsoil. In general, soils are considered to be subsoils. The
Type drop-down menu for the Hypres data set includes Coarse, Medium, Medium fine, Fine, Very fine and
Organic soils.

Note:
Only soil layers that are located not more than 1 m below the ground surface are considered to be Top soils.

The selected soil type and grading (particle fractions) is indicated in the soil texture triangle. As an
alternative, the user can also select the type of soil by clicking one of the sections in the triangle or by
manually specifying the particle fraction values (Figure 153 (on page 231)).

PLAXIS 230 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 153: Groundwater parameters for Hypres data set (Wösten, Lilly, Nemmes & Bas, 1999)

The predefined parameters for both the Van Genuchten model as well as the Approximate Van Genuchten
model are shown in Table 9 (on page 231) and Table 10 (on page 232).

Table 9: Hypres series with Van Genuchten parameters (Wösten, Lilly, Nemmes & Bas, 1999)

Soil class θr (-) θs (-) ksat (m/day) ga (1/m) gl (-) gn (-)

Topsoil

coarse 0.025 0.403 0.600 3.83 1.2500 1.3774

medium 0.010 0.439 0.121 3.14 -2.3421 1.1804

medium fine 0.010 0.430 0.023 0.83 -0.5884 1.2539

fine 0.010 0.520 0.248 3.67 -1.9772 1.1012

very fine 0.010 0.614 0.150 2.65 2.5000 1.1033

Subsoil

coarse 0.025 0.366 0.700 4.30 1.2500 1.5206

medium 0.010 0.392 0.108 2.49 -0.7437 1.1689

PLAXIS 231 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Subsoil

medium fine 0.010 0.412 0.040 0.82 0.5000 1.2179

fine 0.010 0.481 0.085 1.98 -3.7124 1.0861

very fine 0.010 0.538 0.082 1.68 0.0001 1.073

organic 0.010 0.766 0.080 1.30 0.4000 1.2039

Table 10: Hypres series with Approximate Van Genuchten parameters

Soil Class ψs(m) ψk (m)

Topsoil

coarse -2.37 -1.06

medium -4.66 -0.50

medium fine -8.98 -1.20

fine -7.12 -0.50

very fine -8.31 -0.73

Subsoil

coarse -1.82 -1.00

medium -5.60 -0.50

medium fine -10.15 -1.73

fine -11.66 -0.50

very fine -15.06 -0.50

organic -7.35 -0.97

3. USDA:
The USDA series is another international soil classification system (Carsel & Parrish, 1988 (on page 553)).
The hydraulic models available for USDA data set are the Van Genuchten model and the Approximate Van
Genuchten (see Material Models Manual).
The Type drop-down menu for the USDA date set includes Sand, Loamy sand, Sandy loam, Loam, Silt, Silt loam,
Sandy clay loam, Clay loam, Silty clay loam, Sandy clay, Silty clay and Clay. The selected soil type and grading
(particle fractions) are different from the Hypres data sets and can be visualised in the soil texture triangle. As
an alternative, the user can also select the type of soil by clicking one of the sections in the triangle or by
manually specifying the particle fraction values (Figure 154 (on page 233)).

PLAXIS 232 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 154: Groundwater parameters for USDA data set (Carsel & Parrish, 1988)

The parameters for the Van Genuchten and the Approximate Van Genuchten models are shown in Table 11 (on
page 233) and Table 12 (on page 234) .

Table 11: USDA series with Van Genuchten parameters, gl = 0.5 (Carsel & Parrish, 1988)

Soil texture θr (-) θs (-) ksat (m/day) ga (1/m) gn (-)

sand 0.045 0.43 7.13 14.5 2.68

loamy sand 0.057 0.41 3.50 12.4 2.28

sandy loam 0.065 0.41 1.06 7.5 1.89

loam 0.078 0.43 0.25 3.6 1.56

silt 0.034 0.46 0.60 1.6 1.37

silty loam 0.067 0.45 0.108 2.0 1.41

sandy clay loam 0.100 0.39 0.314 5.9 1.48

clayey loam 0.095 0.41 0.624 1.9 1.31

silty clayey loam 0.089 0.43 0.168 1.0 1.23

sandy clay 0.100 0.38 0.288 2.7 1.23

PLAXIS 233 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Soil texture θr (-) θs (-) ksat (m/day) ga (1/m) gn (-)

silty clay 0.070 0.36 0.00475 0.5 1.09

clay 0.068 0.38 0.0475 0.8 1.09

Table 12: USDA series with Approximate Van Genuchten parameters

Soil texture ψs (m) ψk (m)

sand -1.01 -0.50

loamy sand -1.04 -0.50

sandy loam -1.20 -0.50

loam -1.87 -0.60

silt -4.00 -1.22

silty loam -3.18 -1.02

sandy clay loam -1.72 -0.50

clayey loam -4.05 -0.95

silty clayey loam -8.23 -1.48

sandy clay -4.14 -0.55

silty clay -31.95 -0.95

clay -21.42 -0.60

Note:
In Table 9 (on page 231) to Table 11 (on page 233)the symbols θr and θs stand for the residual and the
saturated water content correspondingly. Based on van Genuchten (1980) (on page 554), the effective
degree of saturation is calculated as Seff = (θ-θr)/(θs- θr), in which θ is the water content.
In PLAXIS 2D, the effective degree of saturation is calculated as Seff = (S- Sr) / (Ss-Sr) (see Conventional and
unsaturated soil behaviour in PLAXIS (on page 565)) , in which S is the degree of saturation, Sr the residual
degree of saturation and Ss the saturated degree of saturation. Considering that S = θ / n and under the
assumption that Ss = θs / n = 1, Sr could be derived from Table 9 (on page 231) to Table 11 (on page 233) as Sr
= θ r / θ s.

4. Staring:
The Staring series is a soil classification system which is mainly used in The Netherlands (Wosten, Veerman,
DeGroot & Stolte, 2001 (on page 554)). The hydraulic models available for Staring data set are the Van
Genuchten model and the Approximate Van Genuchten (see the Material Models Manual - Hydraulic Models ).

PLAXIS 234 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 155: Groundwater parameters for Staring data set (Wösten, Veerman, DeGroot & Stolte, 2001)

A distinction can be made between Topsoil and Subsoil. In general, soils are considered to be subsoils. The
Type drop-down menu for the Staring series (Figure 155 (on page 235)) contains the following subsoils
(Wosten, Veerman, DeGroot & Stolte, 2001 (on page 554)):
Non-loamy sand (O1), Loamy sand (O2), Very loamy sand (O3), Extremely loamy sand (O4), Coarse sand (O5),
Boulder clay (O6), River loam (O7), Sandy loam (O8), Silt loam (O9), Clayey loam (O10), Light clay (O11), Heavy
clay (O12), Very heavy clay (O13), Loam (O14), Heavy loam (O15), Oligotrophic peat (O16), Eutrophic peat
(O17) and Peaty layer (O18), and the following topsoils: Non-loamy sand (B1), Loamy sand (B2), Very loamy
sand (B3), Extremely loamy sand (B4), Coarse sand (B5), Boulder clay (B6), Sandy loam (B7), Silt loam (B8),
Clayey loam (B9), Light clay (B10), Heavy clay (B11), Very heavy clay (B12), Loam (B13), Heavy loam (B14),
Peaty sand (B15), Sandy peat (B16), Peaty clay (B17) and Clayey peat (B18).
The selected soil type and grading (particle fractions) are different from the Hypres and the USDA data sets.
The parameters of the hydraulic model for the selected soil type are displayed in the Soil tab at the right side
of the Groundwater tabsheet and are summed up in the tables below.

Note:
Only soil layers that are located not more than 1 m below the ground surface are considered to be Topsoils. In
the tables below the notation "B" corresponds to Topsoils; and the notation "O" indicates Subsoil.

Table 13: Staring series with Van Genuchten parameters (Heinen, Bakker & Wösten, 2001)

Soil class θr (-) θs (-) ksat (m/day) ga (1/m) gl (-) gn (-)

Sand

B1 0.02 0.43 0.234 2.34 0.000 1.801

PLAXIS 235 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Soil class θr (-) θs (-) ksat (m/day) ga (1/m) gl (-) gn (-)

Sand

B2 0.02 0.42 0.125 2.76 -1.060 1.491

B3 0.02 0.46 0.154 1.44 -0.215 1.534

B4 0.02 0.46 0.292 1.56 0.000 1.406

B5 0.01 0.36 0.529 4.52 -0.359 1.933

B6 0.01 0.38 1.01 2.22 -1.747 1.238

O1 0.01 0.36 0.152 2.24 0.000 2.286

O2 0.02 0.38 0.127 2.13 0.168 1.951

O3 0.01 0.34 0.109 1.70 0.000 1.717

O4 0.01 0.35 0.0985 1.55 0.000 1.525

O5 0.01 0.32 0.250 5.21 0.000 2.374

O6 0.01 0.33 0.340 1.62 -1.330 1.311

O7 0.01 0.51 0.391 1.23 -2.023 1.152

Silt

B7 0.00 0.40 0.141 1.94 -0.802 1.250

B8 0.01 0.43 0.0236 0.99 -2.244 1.288

B9 0.00 0.43 0.0154 0.65 -2.161 1.325

O8 0.00 0.47 0.0907 1.36 -0.803 1.342

O9 0.00 0.46 0.0223 0.94 -1.382 1.400

O10 0.01 0.48 0.0212 0.97 -1.879 1.257

Clay

B10 0.01 0.43 0.007 0.64 -3.884 1.210

B11 0.01 0.59 0.0453 1.95 -5.901 1.109

B12 0.01 0.54 0.0537 2.39 -5.681 1.094

O11 0.00 0.42 0.138 1.91 -1.384 1.152

PLAXIS 236 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Clay

O12 0.01 0.56 0.0102 0.95 -4.295 1.158

O13 0.01 0.57 0.0437 1.94 -5.955 1.089

Loam

B13 0.01 0.42 0.130 0.84 -1.497 1.441

B14 0.01 0.42 0.008 0.51 0.000 1.305

O14 0.01 0.38 0.0151 0.30 -0.292 1.729

O15 0.01 0.41 0.037 0.71 0.912 1.298

Peat

B15 0.01 0.53 0.813 2.42 -1.476 1.280

B16 0.01 0.80 0.0679 1.76 -2.259 1.293

B17 0.00 0.72 0.0446 1.80 -0.350 1.140

B18 0.00 0.77 0.0667 1.97 -1.845 1.154

O16 0.00 0.89 0.0107 1.03 -1.411 1.376

O17 0.01 0.86 0.0293 1.23 -1.592 1.276

O18 0.01 0.57 0.0345 1.38 -1.204 1.323

Table 14: Staring series with Approximate Van Genuchten parameters

Class ψs(m) ψk (m)

Sand

B1 -1.87 -1.35

B2 -2.32 -0.79

B3 -3.37 -2.18

B4 -3.81 -2.36

B5 -1.31 -0.56

B6 -4.51 0.70

PLAXIS 237 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Class ψs(m) ψk (m)

Sand

O1 -1.48 -1.15

O2 -1.79 -1.51

O3 -2.46 -1.93

O4 -3.22 -2.30

O5 -1.11 -0.48

O6 -4.50 -1.18

O7 -9.76 -0.98

Silt

B7 -4.72 -1.14

B8 -7.03 -1.46

B9 -9.61 -2.33

O8 -4.78 -1.74

O9 -5.89 -2.07

O10 -7.77 -1.59

Clay

B10 -13.06 -1.47

B11 -9.51 -0.27

B12 -9.61 -0.21

O11 -7.20 -0.76

O12 -11.55 -0.82

O13 -11.45 -0.24

Loam

B13 -6.19 -2.25

B14 -12.82 -7.09

PLAXIS 238 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Loam

B12 -9.61 -0.21

O14 -20.68 -9.37

O15 -9.30 -14.08

Peat

B15 -3.77 -0.73

B16 -4.45 -0.82

B17 -8.03 -1.19

B18 -6.98 -0.64

O16 -5.64 -1.86

O17 -6.06 -1.39

O18 -4.93 -1.45

5. User-defined:
The user-defined option enables the user to define both saturated and unsaturated properties manually.
Please note that this option requires adequate experience with unsaturated groundwater flow modelling. The
hydraulic models available are:

This well-known and widely accepted model requires direct input of the residual
Van Genuchten saturation Sres, the saturation at p = 0, Ssat and the three fitting parameters gn, ga and
gl (see Material Models Manual- Hydraulic models).

The Spline function requires direct input of the capillary height ψ (in unit of length),
the relative permeability Kr (-), and the degree of saturation Sr (-). Data for the Spline
function can be entered by clicking the Table tab. The degree of saturation at
Spline saturated conditions Ssat equals the value assigned to Sr at ψmax (e.g. -ψ = 0 m) and the
residual degree of saturation Sres equals the minimum value assigned to Sr at ψmin (e.g
-ψ = 20 m). During the calculations, the flow calculation kernel employs 'smooth'
relationships based on a spline function between Kr-ψ and Sr-ψ.

When the Saturated option is selected, no extra data input is required. During the
Saturated calculations, PLAXIS 2D will continuously use the saturated permeabilities for soil
layers where a Saturated data set was assigned.

PLAXIS 239 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 156: Groundwater parameters for user-defined data set

6.1.5.2 Saturated permeabilities (kx and ky)


Coefficients of permeability (hydraulic conductivity) have the dimension of velocity (unit of length per unit of
time). The input of permeability parameters is required for groundwater flow and consolidation calculations. In
such calculations, it is necessary to specify the coefficient of permeability for all soil materials with drainage type
Drained, Undrained A and Undrained B, including almost impermeable layers, except for fully impervious layers
with a Non-porous drainage type and layers using a total stress analysis (Undrained C). PLAXIS 2D allows for the
anisotropic permeability of soils where the anisotropy directions coincide with the principal axes x and y.
Users can choose to either enter the values for the permeabilities directly, or choose default values. When opting
for default values there are 2 available options:

From data set The values of kx and ky are obtained from the selected data set.

From grain size The values for kx and ky are obtained from particle size distribution as explained above.
distribution The value for θsat is calculated internally using the equation, θsat = e / (1+e).

In real soils, the difference in permeabilities between the various layers can be quite large. However, care should
be taken when very high and very low permeabilities occur simultaneously in a finite element model, as this
could lead to ill-conditioning of the flow matrix. In order to obtain accurate results, the ratio between the highest
and lowest permeability value in the geometry should not exceed 105.
Note that the input field for permeabilities are greyed out when the drainage type of the material is either Non-
porous or Undrained C.

PLAXIS 240 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

In case of a Standard, Hypres or USDA data set, values for the permeability can be automatically set by either
selecting the From data set option or From grain size distribution.
When a predefined soil data set is chosen or the user specifies the individual particle fractions manually,
assuming the soil has a log-normal function of particle size distribution, then the geometric mean particle
diameter dg and a geometric standard deviation σg can be calculated as:
d g = exp (mcl ln (dcl ) + msi ln (dsi ) + msaln (dsa)) Eq. [17]
1
3 3 2
σg = exp ∑ mi ln (di )2 − ( ∑ mi ln (di )) 2 Eq. [18]
i=1 i=1

where mcl, msi and msa are particle fractions for clay, silt and sand ; dcl, dsi and dsa are the particle size means for
clay, silt and sand respectively (i.e. dcl = 1μm , dsi = 26 μm, dsa = 1.025mm)
According to Aukenthaler, Brinkgreve & Haxaire (2016) (on page 552) The Specific surface area (SSA) can be
approximated using the geometric mean particle diameter in mm:
SSA = 3.89d g −0.905

From the above, the hydraulic conductivity of saturated soils (permeabilities) can be obtained from the soil
texture and porosity of the soil using the following relation:
k x = k y = K sat

K sat = 4 × 10−5 ( 0.5


1 − θsat)1.3b
× exp ( − 6.88mcl − 3.63msi − 0.025) Eq. [19]

b = d g −0.5 + 0.2σg
where
θsat = volumetric water content of saturated soil which is equal to the porosity
of the soil ( n ( n = e / ( 1 + e )) .
dg = the geometric mean particle diameter.
σg = the geometric standard deviation.

6.1.5.3 Void ratio dependency


This advanced feature is to account for the change of permeability with void ratio during a consolidation
analysis. This can be applied by entering a proper value for the ck parameter and the initial void ratio einit. On
entering a real value, the permeability will change according to the formula:

log ( )=
k
k0
Δe
ck
Eq. [20]

where Δe is the change in void ratio, k is the permeability in the calculation and k0 is the input value of the
permeability in the data set (= kx and ky). Note that a proper input of the initial void ratio einit, in the General
tabsheet is required. It is recommended to use a changing permeability only in combination with the Hardening
Soil model, Hardening Soil model with small-strain stiffness, Soft Soil model or the Soft Soil Creep model. In that
case the ck-value is generally in the order of the compression index Cc. For all other models the ck-value should
be left to its default value of 1015.

PLAXIS 241 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

6.1.5.4 Unsaturated zone (ψunsat)


ψunsat (in unit of length relative to the phreatic level) sets the maximum pressure head until which the Mualem-
Van Genuchten functions are used for calculation of relative permeability and degree of saturation. The negative
sign indicates suction. Above the level of ψunsat, the value of Kr and S remains constant. In this way a minimum
degree of saturation (Smin) is guaranteed (Figure 157 (on page 242)). It is used to limit the relative permeability
Kr and degree of saturation for high unsaturated zones.

Figure 157: Relative permeability vs. Degree of saturation

By default a very large value is assigned to ψunsat (= 104). This value is only an indication that the unsaturated
zone is by default unlimited.

6.1.6 Thermal tabsheet [ADV]


The Thermal tabsheet (Figure 158 (on page 243)) contains thermal parameters that are required when dealing
with problems involving (a change of) temperature and the influence on the stress, deformation or groundwater
flow. The meaning of these parameters is described below.

PLAXIS 242 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 158: Thermal tabsheet of the Soil window

6.1.6.1 Thermal diffusion parameters


The first parameters that are required in order to do a thermal flow analysis are the parameters for the thermal
flow or thermal diffusion itself.
The Specific heat capacity of the solid material, cs, is a parameter that describes the amount of energy (heat) that
can be stored in the solid material (i.e. the soil particles) per unit of mass. It is specified in the unit of energy per
unit of mass per unit of temperature. The larger the specific heat, the more energy it takes to increase the
temperature of the material.
The total heat storage in a solid material is the product of the density, ρs, and the specific heat capacity, cs.
Considering soil as a porous medium, the heat storage of the soil is composed of the heat storage in the soil
particles and the heat storage in the pore fluid. The following situations are considered:

For non-porous (ρC)soil = ρsCs


material

For dry material (ρC)soil = (1-n) ρsCs

For phase transition (ρC)soil = (1-n) ρsCs + nS [(1-wu)ρiCi + wu ρwCw] + n (1-S) ρvCv
(e.g. frozen soil)

In other cases (ρC)soil = (1-n) ρsCs + n S ρwCw + n (1-S) ρv Cv

where
n = porosity = [e/(1+e)] (where e = void ratio)

PLAXIS 243 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

S = degree of saturation
wu = property of ice (water in solid state)
index i = unfrozen water content
index w = property of water (water in liquid state)
index v = property of vapour (water in gas state)

Note:
Note that the properties of water in the three different phases are contained in the Constants tabsheet of the
Project properties window ( Project properties (on page 20)).

The Thermal conductivity of the solid material, λs, is a parameter that describes the rate of energy (heat) that can
be transported in the solid material (i.e. the soil particles). It is specified in the unit of power per unit of length
per unit of temperature. The larger the conductivity, the more energy is transported, resulting in a faster
propagation of a change of temperature in the material.
Considering soil as a porous medium, the total heat conductivity of the soil is composed of the heat conductivity
in the soil particles and the heat conductivity in the pore fluid. The following situations are considered:

For non-porous material λsoil = λs

For dry material λsoil = (1-n) λs

For phase transition λsoil = (1-n) λs + n S [(1-wu)λi +wu λw] + n (1-S) λv

In other cases λsoil = (1-n) λs + n S λw + n (1-S) λv

where
λi = thermal conductivity of ice (water in solid state)
λw = thermal conductivity of water (water in liquid state)
λv = thermal conductivity of vapour (water in gas state)
The Density of the solid material, ρs, is the parameter that describes the density of the soil particles, material
expressed in the unit of mass per unit of volume. For example, the density of quartz is 2.65 t/m3 The density
relates to the unit weight of the material (see Project properties (on page 20)), and it contributes to the total
heat storage.

Note:
Note that PLAXIS 2D does NOT check the consistency between the unit weight and the density.

6.1.6.2 Thermal strain parameters


In addition to parameters related to the thermal diffusion, a parameter describing the mechanical effects of a
change of thermal conditions is required.
The Thermal expansion coefficients, α, describe how much the material expands (or elongates) when the
temperature increases. In other words, the thermal expansion coefficient is the (change of) strain per unit of
temperature. If the thermal expansion is isotropic, all coefficients are equal and a single value can be specified
for isotropic expansion, but PLAXIS 2D can also handle anisotropic thermal expansion in x-, y- and z- direction
through the input of 3 linear thermal expansion coefficients.

PLAXIS 244 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

6.1.6.3 Freezing / thawing parameters


The use of the Phase change option may be relevant in the case of permafrost soil, ground freezing and other
situations of frozen soil. In frozen soils, part of the pore water is present in solid state and part of the water may
still be present in liquid state. By activating the phase change option the part of the water in liquid state as a
function of temperature can be defined by means of the Unfrozen water saturation method.
The available options for Unfrozen water saturation method from the drop-down menu are:

User-defined When this option is selected, the user should provide details of the unfrozen water
content, wu as a function of temperature (described below).

From grain size When this option is selected, the unfrozen water content is automatically calculated
distribution using the Specific surface area (SSA)

user-defined water saturation method: After selection of this option, a table should be created in the right-
hand panel, in which the ratio of unfrozen water content over total water content, wu/W0, is defined as a
function of temperature.

Click the Add row button in the toolbar to add a new row in the table.

Click the Insert button to insert a new row before the selected row in the table.

Click the Delete button to delete the selected row in the table.
The values can be defined by clicking the cell in the table and by typing the value. A plot of the unfrozen water
content as a function of temperature is shown below the table (Figure 160 (on page 246)).

Besides defining the values in the table, there is also the possibility to read data from a file using the Open
button in the toolbar.

The current table can be saved using the Save button in the toolbar enabling the usage in other projects
or validating the effect of the modifications in the current project.

The current table can be using the Copy button in the toolbar.

Copied data from other applications (using Ctrl+c) can be imported by using the Paste button. The
Import data window appears (Figure 159 (on page 246). The starting row of the data to be imported can be
defined in the From row cell. The data and the plot is displayed in the Unfrozen water content window after
pressing OK.

Note:
PLAXIS 2D assumes the data file is located in the current project directory when no directory is specified.

PLAXIS 245 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 159: Import data window for Unfrozen water content

Figure 160: A plot for Unfrozen water content as a function of temperature

Clicking the Open .txt file button on the right hand side of the window will open the Open window where the
file can be selected. The file must be an ASCII file that can be created with any text editor. For every line a pair of

PLAXIS 246 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

values (temperature and corresponding unfrozen water content value) must be defined, leaving at least one
space between them. Note that PLAXIS 2D only supports the English notation of decimal numbers using a dot.
In thermal calculations of saturated soils, in which phase transition of the pore water (from water to ice) is
considered, the unfrozen water content is used to identify the amount of water and ice in the pores. This is
important since the different phases of the pore fluid have different thermal and hydraulic properties.
From grain size distribution: When this option is selected, the Unfrozen water content is automatically
calculated using the Specific surface area (SSA) (refer to Eq. [17] ).
SSA = 3.89d g −0.905 Eq. [21]

The temperature dependent unfrozen water content wun can now be determined as follows (Aukenthaler,
Brinkgreve & Haxaire, 2016 (on page 552)) :

wun = exp (0.2618 + 0.5519ln (SSA) − 1.4495SSA−0.2640ln (ΔT )) Eq. [22]

The unfrozen water saturation can now be determined as (Aukenthaler, Brinkgreve & Haxaire, 2016 (on page
552)):
wun ρs
w0
= 100ρw e
wun Eq. [23]

where ρw is the density of pure free water, ρs is the density of solids and e is the void ratio; ΔT is the temperature
difference (positive value) between the bulk freezing point temperature (= 273.16K) and the temperature below
the freezing point and SSA is the specific surface area.

6.1.6.4 Vapour diffusion coefficient


The Vapour diffusion coefficient, Dv, governs the diffusion of vapour in the material. A value of zero disables the
mass flux of vapour in the material.

6.1.6.5 Thermal diffusion enhancement factor


The Thermal diffusion enhancement factor, fTv, influences the dependence of the temperature on the mass flux of
vapour. A value of zero means that the mass flux of vapour is governed only by the variations of pore pressure.

Note:
By default, vapour is not considered, i.e. the Vapour diffusion coefficient and the Thermal diffusion enhancement
factor are set to zero.

6.1.7 Interfaces tabsheet


The properties of interface elements are related to the soil model parameters of the surrounding soil. The
required parameters to derive the interface properties are defined in the Interfaces tabsheet of the Soil
window. These parameters depend on the material model selected to represent the behaviour of the
surrounding soil. In case the Linear Elastic model, the Mohr-Coulomb model, the Hardening Soil model, the
Hardening Soil model with small-strain stiffness, the UBC3D-PLM model, the Soft Soil model, the Soft Soil Creep
model, the Jointed Rock model, the Hoek-Brown model, the NGI-ADP model or the UDCAM-S model has been

PLAXIS 247 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

selected as the Material model, the strength reduction factor Rinter is the main interface parameter (see Figure
161 (on page 248)). In case of the Modified Cam-Clay model, the interface parameters required are the effective
cohesion c'ref, the effective friction angle φ' and the dilatancy angle ψ. In case of the User-defined Soil Model, the
tangent stiffness for primary oedometer loading Eoedref, the effective cohesion c'ref, the effective friction angle φ',
the dilatancy angle ψ and the parameters UD-Power and UD-Pref are required as interface parameters. For more
information on the interface parameters required for the User-defined Soil Model (see the Material Models
Manual - Input of UD model parameters via user-defined.

Figure 161: Interfaces tabsheet of the Soil window

6.1.7.1 Stiffness
With the option Stiffness determination the user can choose to either have the interface stiffness Derived from the
soil parameters, or to have a Direct input of the interface stiffness.
Derived: by default, stiffness is set to Derived. The shear and compression moduli are related by the
expressions:
1 − νi
Eoed ,i = 2Gi 1 − 2ν
i
2 Eq. [24]
Gi = RinterGsoil ≤ Gsoil
νi = 0.45

When using user-defined material model, the user inserts the interface stiffness in the form of a power law
formulation with Eoedref along with UD-pref and UD-Power.
Direct: this option allows to directly specify the kn and ks values.

PLAXIS 248 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

kn Elastic interface normal stiffness kN/m3

ks Elastic interface shear stiffness kN/m3

6.1.7.2 Strength
In case of the Linear Elastic model, the Mohr-Coulomb model, the Hardening Soil model, the Hardening Soil
model with small-strain stiffness, the UBC3D-PLM model, the Soft Soil model, the Soft Soil Creep model, the
Jointed Rock model, the Hoek-Brown model, the NGI-ADP model or the UDCAM-S model, the interface strength is
defined by the parameter Rinter. The interface strength can be set using the following options:
Rigid: This option is used when the interface should not have a reduced strength with respect to the strength in
the surrounding soil. For example, extended interfaces around corners of structural objects (Figure 68 (on page
108)) are not intended for soil-structure interaction and should not have reduced strength properties. The
strength of these interfaces should be assigned as Rigid (which corresponds to Rinter = 1.0). As a result, the
interface properties, including the dilatancy angle ψi, are the same as the soil properties in the data set, except
for Poisson's ratio νi (see further).
Manual: The value of Rinter can be entered manually if the interface strength is set to Manual. In general, for real
soil-structure interaction the interface is weaker and more flexible than the surrounding soil, which means that
the value of Rinter should be less than 1. Suitable values for Rinter for the case of the interaction between various
types of soil and structures in the soil can be found in the literature. In the absence of detailed information it may
be assumed that Rinter is of the order of 2/3. A value of Rinter greater than 1 cannot be used.
When the interface is elastic then both slipping (relative movement parallel to the interface) and gapping or
overlapping (i.e. relative displacements perpendicular to the interface) could be expected to occur.
The magnitudes of the interface displacements are:
σn σn ti
Elastic gap displacement = kn
= Eoed ,i
Eq. [25]

τti
τ
Elastic slip displacement = ks
= Gi
Eq. [26]

where Gi is the shear modulus of the interface, Eoed,i is the one-dimensional compression modulus of the
interface, ti is the virtual thickness of the interface generated during the creation of interfaces in the geometry
model (Interfaces (on page 104)), kn is the elastic interface normal stiffness and ks is the elastic interface shear
stiffness.

Note:
Note that a reduced value of Rinter not only reduces the interface strength, but also the interface stiffness.

It is clear from these equations that, if the elastic parameters are set to low values, the elastic displacements may
be excessively large. If the values of the elastic parameters are too large, however, this can result in numerical ill-
conditioning of the stiffness matrix. The key factor in the stiffness is the virtual thickness. This value is
automatically chosen such that an adequate stiffness is obtained. The user may change the virtual thickness. This
can be done in the Interface window that appears after double clicking an interface in the geometry model
(Interfaces (on page 104)).
Manual with residual strength: When the limit value of the interface strength as defined by Rinter is reached,
the interface strength may soften down to a reduced value as defined by Rinter,residual. Definition of the Rinter,residual
is possible when the Manual with residual strength option is selected for the interface strength.

PLAXIS 249 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Interface reduction factor (Rinter): An elastic-plastic model is used to describe the behaviour of interfaces for
the modelling of soil-structure interaction. The Coulomb criterion is used to distinguish between elastic
behaviour, where small displacements can occur within the interface, and plastic interface behaviour when
permanent slip may occur. For the interface to remain elastic the shear stress τ is given by:
| τ | < − σn tan (φi ) + ci Eq. [27]

where σn is the effective normal stress.


For plastic behaviour τ is given by:
| τ | = − σn tan (φi ) + ci Eq. [28]

where φi and ci are the friction angle and cohesion (adhesion) of the interface. The strength properties of
interfaces are linked to the strength properties of a soil layer. Each data set has an associated reduction factor
for interfaces Rinter. The interface properties are calculated from the soil properties in the associated data set and
the strength reduction factor by applying the following rules:
ci = Rintercsoil
Eq. [29]
tan (φi ) = Rintertan (φsoil ) ≤ tan (φsoil )

ψi = 0 for Rinter < 1, otherwise ψi = ψsoil Eq. [30]

In the Hardening Soil model, the Hardening Soil model with small-strain stiffness and the UBC3D-PLM model,
φsoil is always referred to the failure-peak value of φ chosen as input parameter.
In addition to Mohr-Coulomb's shear stress criterion, the tension cut-off criterion, as described before (see
Mechanical tabsheet (on page 172)), also applies to interfaces (if not deactivated):
σn < σt,i = Rinterσt ,soil Eq. [31]

where σt,soil is the tensile strength of the soil.


Interface strength in undrained condition: When the drainage condition is Undrained B or Undrained C, the
calculation is performed with undrained strength parameters. For the interface to remain elastic the shear stress
τ is given by:
| τ | < su Eq. [32]

where su is the undrained shear strength.


For plastic behaviour τ is given by:
| τ | = su Eq. [33]

The interface property is calculated from the soil property in the associated data set and the strength reduction
factor Rinter by applying the following rule:
su,i = Rintersu,soil Eq. [34]

when su,i is the undrained shear strength of the interface and su,soil is the undrained shear strength of the soil.
The tension cut-off criterion applies as in calculation with effective parameters for strength.
In the NGI-ADP model, su,soil is computed accordingly to what is reported in the Material Models Manual.
Residual interface strength (Rinter, residual): When the Manual with residual strength option is selected the
parameter Rinter, residual can be specified. The interface strength will reduce to the residual strength as defined by
(Rinter, residual) and the strength properties of the soil, as soon as the interface strength is reached.

PLAXIS 250 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Note:
• Note that the same values of partial factors in Design approaches are applied to both interface strength Rinter
and residual interface strength Rinter, residual.
• A reduced residual strength is not recommended to be used in Safety calculations.

Consider gap closure: When the interface tensile strength is reached a gap may occur between the structure
and the soil. When the load is reversed, the contact between the structure and the soil needs to be restored
before a compressive stress can developed. This is achieved by selecting the Consider gap closure option in the
Interfaces tabsheet of the Soil window. If the option is NOT selected, contact stresses will immediately develop
upon load reversal, which may not be realistic.
Interfaces using the Hoek & Brown model : When using the Hoek-Brown model as a continuum model to
describe the behaviour of a rock section in which interface elements are used, equivalent interface strength
properties φi, ci and σt,i are derived from this model. The general shear strength criterion for interfaces as well
as the tensile strength criterion are still used in this case:
| τ | ≤ − σn tan (φi ) + ci Eq. [35]

σn ≤ σt,i Eq. [36]

Starting point for the calculation of the interface strength properties is the minor principal effective stress σ'3 in
the adjacent continuum element. At this value of confining stress the tangent to the Hoek-Brown contour is
calculated and expressed in terms of φ and c:
f̄ ′
sin(φ) = 2 + f̄ ′
Eq. [37]
c=
1 − sin(φ )
2cos (φ )
( f̄ + 2σ ′ 3sin (φ )
1 − sin (φ )
)
where

(
f̄ = σci mb
−σ ′3
σci
+s ) a

Eq. [38]
f̄ ′ = (−σ ′3
amb mb σ
ci
+s ) a−1

and a, mb, s and σci are the Hoek-Brown model parameters in the corresponding material data set. The interface
friction angle φ'i and adhesion c'i as well as the interface tensile strength σt,i are now calculated using the
interface strength reduction factor Rinter:
tan (φi ) = Rintertan (φ )
ci = = Rinterc
Eq. [39]
sσci
σt,i = Rinterσt = Rinter m
b

For more information about the Hoek-Brown model and an explanation of its parameters, reference is made to
the Material Models Manual.
Interfaces using the Modified Cam-Clay model If the Modified Cam-Clay model is selected in the General
tabsheet to describe the behaviour of the surrounding soil, the following parameters are required to model the
interface behaviour:

PLAXIS 251 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

cref Cohesion of the interface [kN/m2]

φ'i Internal friction angle of the interface [°]

ψi Dilatancy angle of the interface [°]

When the interface is elastic then both slipping (relative movement parallel to the interface) and gapping or
overlapping (i.e. relative displacements perpendicular to the interface) could be expected to occur.
The magnitudes of these displacements are:
σn σn ti
Elastic gap displacement = kn
= Eoed ,i
Eq. [40]

τ τti
Elastic slip displacement = ks
= Gi
Eq. [41]

where Gi is the shear modulus of the interface, Eoed,i is the one-dimensional compression modulus of the
interface and ti is the virtual thickness of the interface, generated during the creation of interfaces in the
geometry model (Interfaces (on page 104)). kn is the elastic interface normal stiffness and ks is the elastic
interface shear stiffness. The shear and compression moduli are related by the expressions:

3 (1 − νi ) σn
Eoed ,i = λ (1 + νi ) (1 + e0)

3(1 − 2νi ) σn Eq. [42]


Gi = 2(1 + νi ) λ (1 + e0)

νi = 0.45

6.1.7.3 Real interface thickness (δinter )


The real interface thickness δinter is a parameter that represents the real thickness of a shear zone between a
structure and the soil. The value of δinter is only of importance when interfaces are used in combination with the
Hardening Soil model or the Hardening Soil model with small-strain stiffness. The real interface thickness is
expressed in the unit of length and is generally of the order of a few times the average grain size. This parameter
is used to calculate the change in void ratio in interfaces for the dilatancy cut-off option. The dilatancy cut-off in
interfaces can be of importance, for example, to calculate the correct bearing capacity of tension piles.

6.1.7.4 Groundwater flow through interfaces


In situations involving groundwater flow, consolidation or fully coupled flow-deformation analysis, interface
elements can contribute to the flow of groundwater and thereby influence the pore pressure distribution.
Therefore, interface permeabilities are relevant in such situations. It is also important for the modelling of flow
through a system of joints or faults in low permeable rocks when analysing groundwater flow, pore pressures
and seepage for applications in rock.
Flow in interface elements may involve flow through the element, as well as, flow in the longitudinal direction of
the element (see Figure 163 (on page 255)). Therefore, the following distinction is made:

PLAXIS 252 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

1. Cross permeability, defined by means of the Hydraulic resistance, where water can flow through the
interface.
The cross permeability can be set using the following options (see Figure 162 (on page 254)):
• Impermeable [No user input required]: This is the default option. This option sets the interface to have a
zero cross permeability (infinite cross resistance), provided the interface is active in flow as defined by the
interface settings in the Selection explorer (see Table 3 (on page 107)). Note that there is no output of
groundwater flow qn available.
• Semi-permeable: In this case the interface has a specific non-zero hydraulic resistance, expressed in the
unit of time (hydraulic delay), provided the interface is active in flow as defined by the interface settings in
the Selection explorer (see Table 3 (on page 107)). The output program will show the amount of
groundwater flow qn crossing the interface.
Hydraulic resistance (d/k): The hydraulic resistance defines the permeability through the interface
(perpendicular to the interface longitudinal direction). Considering a semi-permeable wall with a
thickness d and permeability (or hydraulic conductivity) k, the hydraulic resistance is defined by d/k,
expressed in the unit of time.To determine d/k, one needs to measure the average discharge q through a
wall (per unit of area) for a given head difference Δh, so d/k = Δh/q.
• Fully permeable [No user input required]: In this case the interface has an infinite cross permeability (zero
cross resistance), irrespective whether it is active in flow as defined by the interface settings in the
Selection explorer (see Table 3 (on page 107)). Note that there is no output of groundwater flow
available when choosing this option.

PLAXIS 253 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 162: Cross permeability options for interfaces

PLAXIS 254 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

2. Drainage conductivity [dk], where the interface acts as an additional flow "channel". The drainage
conductivity defines the permeability in the interface longitudinal direction parallel to the local axes
(drainage capacity), provided the interface is active in flow as defined by the interface settings in the
Selection explorer (see Table 3 (on page 107)). The drainage conductivity is expressed in the unit of volume
(water volume that is transported in the interface longitudinal direction) per unit of time per unit of width in
the out-of-plane direction. The default value is zero, which means that there is no drainage capacity in the
interface.
Considering a semi-permeable gap with a thickness d and permeability k between two impermeable media,
the drainage conductivity is defined by the product of d and k (dk), expressed in the unit of volume per unit of
time per unit of width in the out-of-plane direction. This quantity defines the total amount of water that is
transported through the gap (drain) per unit of time per unit width. In order to determine dk, one needs to
measure the total discharge Q/b through the gap (per unit of width in the out-of-plane direction) for a given
head difference Δh and a given length of the gap L, such that dk = (Q/b) (L/Δh).

Figure 163: Cross permeability(left) and Drainage conductivity (right)

Note: Notice that, neither for the hydraulic resistance nor for the drainage conductivity, the actual thickness d
and permeability k really matter.

Practical remarks: The interface elements are generally at both sides of a wall: for example, in the case of an
excavation where the wall is to some extent permeable. In such a case it is suggested to assign the appropriate
interface permeability to the 'outside' (soil side) of the wall, whereas the interface at the 'inside' (side that is
excavated) is made fully permeable, even along the embedded part of the wall below the excavation. Similarly
for a semi-permeable tunnel lining: the appropriate interface permeability should be given to the interface at the
outside of the tunnel lining, whereas the interface at the inside (which will be excavated) should be fully
permeable. Hence, this requires an additional material data set to be defined which should be assigned directly
to the interface.

6.1.7.5 [ADV] Thermal resistance


In Thermal calculations, interface elements can act as insulators, separating the temperature at either side of the
interface. As interfaces behave in groundwater flow calculations (where they block the flow completely,
behaving as semi-permeable or do not form any resistance at all), interfaces, also, transfer some heat in thermal
calculations. The Thermal resistance, Rthermal, is used to control the heat transfer across the interface. Thermal
resistance is defined as the ratio of the interface thickness, d, and the thermal conductivity of the interface
material, λ:
d
Rthermal = λ
Eq. [43]

PLAXIS 255 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Thermal resistance values are generally provided by manufacturers of insulating materials that are produced as
plates or blankets (foams, wools). Otherwise, the thermal resistance can easily be calculated from the above
equation by considering the material's thermal conductivity and a characteristic thickness of the interface layer.

6.1.7.6 Interfaces below or around corners of structures


When interfaces are extended below or around corners of structures to avoid stress oscillations (Interfaces (on
page 104)), these extended interfaces are not meant to model soil-structure interaction behaviour, but just to
allow for sufficient flexibility. Hence, when using Rinter< 1 for these interface elements an unrealistic strength
reduction is introduced in the ground, which may lead to unrealistic soil behaviour or even failure. Therefore it
is advised to create a separate data set with Rinter = 1 and to assign this data set only to these particular interface
elements. This can be assigned by dropping the appropriate data set on the individual interfaces rather than
dropping it on the associated soil cluster (the interface lines should blink red; the associated soil cluster may not
change colour). Alternatively, you can click the right-hand mouse button on these particular interface elements
and select Properties and subsequently Positive interface element or Negative interface element. In the Interface
window, select the appropriate material set in the Material set drop-down menu and click the OK button.

6.1.7.7 Interface permeability


The permeability of the interfaces can be specified by checking the corresponding check box Active in flow in the
Interface subtree. Impermeable interfaces may be used to block the flow perpendicular to the interface (cross-
permeability) in a consolidation analysis or a groundwater flow calculation, for example to simulate the
presence of an impermeable wall. This is achieved by a full separation of the pore pressure degrees-of-freedom
of the interface node pairs. On the other hand, if interfaces are present in the mesh it may be the user's intension
to explicitly avoid any influence of the interface on the flow and the distribution of (excess) pore pressures, for
example in interfaces around corner points of structures (Interfaces (on page 104)). In such a case the interface
should be semi-permeable or fully permeable. For fully permeable interfaces the pore pressure degrees-of-
freedom of the interface node pairs are fully coupled.

6.1.8 Initial tabsheet


The Initial tabsheet contains parameters to generate the initial stresses by means of the K0 procedure (Figure
164 (on page 257)).

PLAXIS 256 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 164: Soil window (Initial tabsheet of the Mohr-Coulomb model)

The K0-values can be defined automatically by selecting the option Automatic in the K0 determination drop-down
box or manually by selecting the option Manual.

6.1.8.1 K0-values
In general, two K0-values can be specified, one for the x-direction (in-plane) and one for the z-direction (out-of-
plane):
K 0,x = σ ′ xx / σ ′ yy K 0,z = σ ′ zz / σ ′ yy Eq. [44]
The checkbox can be used to set the K0,z value equal to the K0,x value.
The default K0-values are then in principle based on Jaky's formula:
K 0 = 1 − sin (φ ) Eq. [45]

For advanced models (Hardening Soil model, Hardening Soil model with small-strain stiffness, Soft Soil model,
Soft Soil Creep model, Modified Cam-Clay model) the default value is based on the K0nc model parameter and is
also influenced by the OCR-value and POP-value in the following way:
νur
K 0nc POP − POP
νur 1 − νur
K 0,x = K 0,z = K 0nc OCR − (OCR − 1) + Eq. [46]
1 − νur
| σ ′ 0yy |
The POP-value will result in a stress-dependent K0 value within the layers and thus the K0-values are hidden
when automatic K0 determination is chosen (e.g. Figure 165 (on page 258)).

PLAXIS 257 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Figure 165: Soil window (Initial tabsheet of the Hardening Soil model

Be careful with very low or very high K0-values, since these values might bring the initial stress in a state of
failure. For a cohesionless material it can easily be shown that to avoid failure, the value of K0 is bounded by:
1 − sin (φ ) 1 + sin (φ )
1 + sin (φ )
< K0 < 1 − sin (φ )
Eq. [47]

6.1.8.2 OCR and POP


When using advanced models (Hardening Soil model, Hardening Soil model with small-strain stiffness, Soft Soil
model, Soft Soil Creep model, Modified Cam-Clay model, Sekiguchi-Ohta model) an initial preconsolidation stress
has to be determined. In the engineering practice it is common to use a vertical preconsolidation stress, σp, but
PLAXIS 2D needs an equivalent isotropic preconsolidation stress, ppeq to determine the initial position of a cap-
type yield surface. If a material is overconsolidated, information is required about the overconsolidation Ratio
(OCR), i.e. the ratio of the greatest effective vertical stress previously reached, σp (see Figure 166 (on page
259)), and the in-situ effective vertical stress, σ'yy0.
σp
OCR = 0
Eq. [48]
σ ′
yy

PLAXIS 258 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling soil and interface behaviour

Using OCR Using POP


Figure 166: Illustration of vertical preconsolidation stress in relation to the in-situ vertical effective stress

It is also possible to specify the initial stress state using the Pre-Overburden Pressure (POP) as an alternative to
prescribing the overconsolidation ratio. The Pre-Overburden Pressure is defined by:

POP = | σ p − σ ′ 0yy | Eq. [49]

These two ways of specifying the vertical preconsolidation stress are illustrated in Figure 166 (on page 259).
The preconsolidation stress σp is used to compute the equivalent preconsolidation pressure ppeq which
determines the initial position of a cap-type yield surface in the advanced soil models. The calculation of ppeq is
based on the principal stress history (σ'1,max, σ'2, σ'3). The actual determination of ppeq depends on the
constitutive model being used.
The principal stress history is initialised in the Initial phase (K0-procedure or Gravity loading) based on the
Cartesian effective stress components and the pre-overburden pressure (POP) or overconsolidation ratio (OCR)
defined in the boreholes or data set. From this, Cartesian preconsolidation stress levels are calculated based on
the following equations:

σ′ xx,c = K 0nc OCRσ ′ yy


σ′ yy,c = OCRσ ′ yy
Eq. [50]
σ′ zz,c = K 0nc OCRσ ′ yy
σ′ xy,c =σ′ xy

where K0nc is the K0-value associated with normally consolidated states of stress. K0nc is a model parameter in
advanced constitutive models and estimated in simple models (K0nc= 1-sinφ). Models that do not have φ as input
parameter use K0nc=0.5 (φ=30°).
The Cartesian stress components (σ'xx,c, σ'yy,c, σ'zz,c, σ'xy,c) are transformed to principal stress components
(σ'1,max, σ'2, σ'3) and the maximum major principal stress, σ'1,max, is kept as a general state parameter which is
available for succeeding phases. In subsequent phases, σ'1,max is updated if the major principal stress is larger
than the current one.
If, in later calculation phases, the soil behaviour is changed to an(other) advanced material model, the equivalent
preconsolidation pressure ppeq is initialised according to the current (updated) principal stress history (σ'1,max,
σ'2, σ'3).

PLAXIS 259 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling undrained behaviour

Table 15: K0 behaviour

Material Model K0

Linear Elastic model 0.5

Mohr-Coulomb model 1- sin φ

Hardening Soil model νur


K 0nc OCR − 1 − νur
(OCR − 1)

Hardening Soil model with small-strain stiffness νur


K 0nc OCR − 1 − νur
(OCR − 1)

UBC3D-PLM model 1- sin φp

Soft Soil model νur


K 0nc OCR − 1 − νur
(OCR − 1)

Soft Soil Creep model νur


K 0nc OCR − 1 − νur
(OCR − 1)

Jointed Rock model 0.5

Modified Cam-Clay model νur


K 0nc OCR − 1 − νur
(OCR − 1)

NGI-ADP model Calculated by the kernel

UDCAM-S model Calculated by the kernel

Hoek-Brown model 0.5

Sekiguchi-Ohta model (Inviscid) νur


K 0nc OCR − 1 − νur
(OCR − 1)

Sekiguchi-Ohta model (Viscid) νur


K 0nc OCR − 1 − νur
(OCR − 1)

Concrete model 1- sin φmax

User-defined Soil Model 0.5

PLAXIS 260 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling undrained behaviour

6.2 Modelling undrained behaviour


In undrained conditions, no water movement takes place. As a result, excess pore pressures are built up.
Undrained analysis is appropriate when:
• Permeability is low or rate of loading is high.
• Short term behaviour has to be assessed.
Different modelling schemes are possible in PLAXIS 2D to model undrained soil response in a Plastic calculation,
a Safety analysis or a Dynamic analysis. The modelling scheme depends on the selection of the Drainage type
parameter (General tabsheet (on page 165)). More details about these methods are in the Material Models
Manual - Preliminaries on material modelling.

Note:
The Drainage type setting is only considered in a Plastic calculation, a Safety analysis or a Dynamic analysis.
When a Consolidation analysis, a Fully coupled flow-deformation analysis or a Dynamic with consolidation analysis
is performed, the Drainage type is ignored and the soil response is determined by the saturated permeability of
the material that is specified in the Groundwater tabsheet of the material datase.

Note:
The modelling of undrained soil behaviour is even more complicated than the modelling of drained behaviour.
Therefore, the user is advised to take the utmost care with the modelling of undrained soil behaviour.

Before considering the consequences of a particular selection of the drainage type parameter for undrained soil
behaviour, first a general description is given of the various modelling possibilities:
1. Undrained effective stress analysis with effective stiffness parameters
A change in total mean stress in an undrained material during a Plastic calculation phase gives rise to excess
pore pressures. PLAXIS 2D differentiates between steady-state pore pressures and excess pore pressures, the
latter generated due to small volumetric strain occurring during plastic calculations and assuming a low (but
non-zero) compressibility of the pore water. This enables the determination of effective stresses during
undrained plastic calculations and allows undrained calculations to be performed with effective stiffness
parameters. This option to model undrained material behaviour based on effective stiffness parameters is
available for all material models in the PLAXIS 2D. The undrained calculations can be executed with effective
stiffness parameters, with explicit distinction between effective stresses and (excess) pore pressures.
2. Undrained effective stress analysis with effective strength parameters
Undrained effective stress analysis can be used in combination with effective strength parameters φ' and c' to
model the material's undrained shear strength. In this case, the development of the pore pressure plays a
crucial role in providing the right effective stress path that leads to failure at a realistic value of undrained
shear strength (cu or su). However, note that most soil models are not capable of providing the right effective
stress path in undrained loading. As a result, they will produce the wrong undrained shear strength if the
material strength has been specified on the basis of effective strength parameters. Another problem is that
for undrained materials effective strength parameters are usually not available from soil investigation data.
The advantage of using effective strength parameters in undrained loading conditions is that after
consolidation a qualitatively increased shear strength is obtained, although this increased shear strength
could also be quantitatively wrong, for the same reason as explained before.
3. Undrained effective stress analysis with undrained strength parameters

PLAXIS 261 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling undrained behaviour

Especially for soft soils, effective strength parameters are not always available, and one has to deal with
measured undrained shear strength (cu or su) as obtained from undrained tests. Undrained shear strength,
however, cannot easily be used to determine the effective strength parameters φ' and c'. Moreover, even if
one would have proper effective strength parameters, care has to be taken as to whether these effective
strength parameters will provide the correct undrained shear strength in the analysis. This is because the
effective stress path that is followed in an undrained analysis may not be the same as in reality, due to the
limitations of the applied soil model.
In order to enable a direct control on the shear strength, PLAXIS 2D allows for an undrained effective stress
analysis with direct input of the undrained shear strength (Undrained B).

6.2.1 Undrained A
The Drainage type Undrained A enables modelling undrained behaviour using effective parameters for stiffness
and strength. The characteristic features of method Undrained A are:
• The undrained calculation is performed as an effective stress analysis. Effective stiffness and effective
strength parameters are used.
• Pore pressures are generated, but may be inaccurate, depending on the selected model and parameters.
• Undrained shear strength su is not an input parameter but an outcome of the constitutive model. The
resulting mobilised shear strength must be checked against known data.
• Consolidation analysis can be performed after the undrained calculation, which affect the shear strength.
Undrained A drainage type is available for the following models: Linear Elastic model, Mohr-Coulomb model,
Hardening Soil model, Hardening Soil model with small-strain stiffness, UBC3D-PLM model, Soft Soil model, Soft
Soil Creep model, Modified Cam-Clay model and UDSM.

6.2.2 Undrained B
The Drainage type Undrained B enables modelling undrained behaviour using effective parameters for stiffness
and undrained strength parameters. The characteristic features of method Undrained B are:
• The undrained calculation is performed as an effective stress analysis.
• Effective stiffness parameters and undrained strength parameters are used.
• Pore pressures are generated, but may be highly inaccurate.
• Undrained shear strength su is an input parameter.
• Consolidation analysis should not be performed after the undrained calculation. If consolidation analysis is
performed anyway, su must be updated.
Undrained B drainage type is available for the following models: Mohr-Coulomb model, Hardening Soil model,
Hardening Soil model with small-strain stiffness, and UDCAM-S model. Note that when using Undrained B in the
Hardening Soil model or Hardening Soil model with small-strain stiffness, the stiffness moduli in the model are
no longer stress-dependent and the model exhibits no compression hardening.

PLAXIS 262 PLAXIS 2D-Reference Manual


Material properties and material database
Modelling undrained behaviour

6.2.3 Undrained C
The Drainage type Undrained C enables simulation of undrained behaviour using a total stress analysis with
undrained parameters. In that case, stiffness is modelled using an undrained Young's modulus Eu and an
undrained Poisson ratio νu, and strength is modelled using an undrained shear strength cu (su) and φ = φu = 0°.
Typically, for the undrained Poisson ratio a value close to 0.5 is selected (between 0.495 and 0.499). A value of
exactly 0.5 is not possible, since this would lead to singularity of the stiffness matrix. The disadvantage of this
approach is that no distinction is made between effective stresses and pore pressures. Hence, all output
referring to effective stresses should now be interpreted as total stresses and all pore pressures are equal to
zero. Note that a direct input of undrained shear strength does not automatically give the increase of shear
strength with consolidation. The characteristic features of method Undrained C are:
• The undrained calculation is performed as a total stress analysis.
• Undrained stiffness parameters and undrained strength parameters are used.
• Pore pressures are not generated.
• Undrained shear strength su is an input parameter.
• Consolidation analysis should not be performed after the undrained calculation. If consolidation analysis is
performed anyway, su must be updated.
Undrained C drainage type is available for the following models: Linear Elastic model, Mohr-Coulomb model,
NGI-ADP model and UDCAM-S model.

Note:
For Undrained B and Undrained C an increased shear strength with depth can be modelled using the advanced
parameter su,inc.

6.2.4 Undrained behaviour for unsaturated soil


As aforementioned, the build-up of pore pressures plays a crucial role in undrained analyses, especially
Undrained A and Undrained B. Excess pore pressures are generated assuming very low compressibility of pore
fluids. This is only valid for fully saturated soils where voids are filled by water. In unsaturated soils, the porous
space is filled by both water and air. It is thus expected that the compressibility of pore fluid is much higher
compared to saturated soils, and thereby a lower excess pore pressure.
On physical grounds, water in unsaturated soils co-exists with air under the form of water meniscus due to the
capillary effect (surface tension). The excess pore pressure is governed by three factors: the compressibility of
water, compressibility of air and surface tension (suction) effect. The latter is related to the Soil-Water retention
curve (defined in Groundwater tabsheet (on page 229)).
In PLAXIS, we assume that the air pressure is always constant, namely excess air pressure and air
compressibility are not considered (see Appendix Definition of stresses and related quantities (on page 565) ).
With this in mind, the undrained analysis only means that the water phase is undrained, while air phase is
always drained. Excess pore pressure in undrained behaviour for an unsaturated zone is generated following
two cases:
• If Ignore suction (on page 566) is not chosen , the suction effect is considered. Excess pore pressure is
governed by an ''unsaturated bulk modulus of pore fluid'' which takes into account both the compressibility

PLAXIS 263 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for discontinuities

of water and Soil-Water retention curve (see Table 22 (on page 573)). This fluid bulk modulus usually is
much smaller than the water bulk modulus of water.
• If Ignore suction is chosen, the suction effect is neglected, and soil behaves like a saturated soil. Excess pore
pressure is only governed by the compressibility of water in a similar way as saturated soil and is generally
much higher than the case of considering suction. This might probably result in a stiffer undrained volume
change behaviour, but a lower shear strength due to a higher pore pressure build-up. This is certainly
dependent on the constitutive model used.

Note:
The behaviour of unsaturated soil is more realistic but also more complicated than saturated soil. The user is
advised to take the utmost care with the modelling as well as Soil-Water retention curve input parameters.

6.3 Material data sets for discontinuities


As mentioned in the chapter Modelling loads and structures - Structures mode - Discontinuities (on page
113), the material properties for this element can be entered in the material data set window.

6.3.1 General properties

Several data sets may be created to distinguish different discontinuities. As displayed in Figure 167 (on page
265) , the material data set includes the following options:

PLAXIS 264 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for discontinuities

Figure 167: General tabsheet - Material properties of discontinuities

A user may specify any identification title for a discontinuity data set. It is advisable to
Identification use a meaningful name since the data set will appear in the database tree view by its
identification.

Material Model1 Material model to be considered in the discontinuity.

Two options allowed Drained and Non-porous. Discontinuities can be open or filled with
certain porous materials (for instance quartz, calcite or sand) which can be considered
Drainage type as Drained. However, if the infill is considered non-porous (for instance rock grouting
for impermebilisation techniques), the discontinuity element can also be considered
Non-porous.

Colour can be used as a distinction tool in the model. The default color for a
Colour
discontinuity is orange.

Comments A user may write down comments related to the material data set.

Note:
• 1Currently only the Mohr-Coulomb material model is available for the modelling of discontinuity elements.

6.3.2 Mechanical properties

PLAXIS 265 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for discontinuities

The required parameters for the strength and stiffness of discontinuity are defined in the Mechanical tabsheet
of the material properties definition as shown in Figure 168 (on page 266).

Figure 168: Mechanical properties of discontinuities with peak strength

6.3.2.1 Stiffness
Two distinct material stiffness can be assigned to discontinuities kn and ks.

kn Normal stiffness kN/m3

ks Shear stiffness kN/m3

6.3.2.2 Strength
Defines the strength parameters of the material in the discontinuity according the material model available (see
General properties (on page 264)). Two strength methods are available: Peak and Peak and Residual.
Peak
This is the default method. The strength parameters represent the peak strength and no material softening
occurs when the peak strength is reached.

c, Effective cohesion of the material in the discontinuity kN/m2

φ' Effective Internal friction angle of the material in the discontinuity °

PLAXIS 266 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for discontinuities

ψi Dilatancy angle of the material in the discontinuity °

Consider When tensile stresses are applied to a discontinuity, the discontinuity will open, -
gap closure creating a gap in the element. When the load is reversed, the contact between
either sides of the discontinuities needs to be restored before a compressive stress
can develop

Peak and Residual


When this Strength method is selected, the RSF (Residual Strength Factor) can be specified for the discontinuity
(as shown in Figure 169 (on page 267)). Once the peak strength is reached, the discontinuity material strength
will reduce to its residual strength as defined by RSF, as follows:
If RSF < 1 :
▸ cres = RSF ⋅ cini
▸ tan(φ res ) = RSF ⋅ tan(φ ini )

RSF Residual Strength Factor -

Figure 169: Mechanical properties of discontinuities with peak and residual strength

PLAXIS 267 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for discontinuities

6.3.3 Groundwater properties

Groundwater flow parameters are required when dealing with the hydraulic conditions of discontinuities
specifically when drained behaviour is addressed. For this reason, to enable the use of groundwater hydraulic
parameters the Drainage type of material needs to be set as Drained (see General properties (on page 264)).
When drained conditions are selected, the Groundwater tabsheet is displayed as in Figure 170 (on page 268).

Figure 170: Groundwater properties of discontinuities

The groundwater parameter available is listed below:

T1 Transmissivity in the local direction 1 of a discontinuity: volumetric flow rate m3/day/m


though the discontinuity in this local direction per unit width (perpendicular to the
flow direction) under unit hydraulic gradient

It is important to mention that under drained conditions the discontinuity behaves as Fully permeable (the
discontinuity has an infinite cross permeability or zero cross resistance), and the Transmissivity represents the
drainage conductivity in the longitudinal direction of the fracture. .

Note: If Drainage type defined in General properties (on page 264) is set as Non-porous the Transmissivity (T) is
disabled.

PLAXIS 268 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for plates

6.3.4 Thermal properties [ADV]

The Thermal tabsheet (see Figure 171 (on page 269)) contains discontinuity properties for problems involving
(a change of) temperature and the influence on stress, deformation or groundwater.

Figure 171: Thermal tabsheet - Discontinuity element

The thermal parameter available for discontinuities is listed below:

R Thermal Resistance m2K/kW

The thermal resistance R is used to control the heat transfer across the discontinuity. R is defined as the ratio
between the aperture of the discontinuity, d, and the thermal conductivity, λ, of the discontinuity material.
d
R = λ
Eq. [51]

PLAXIS 269 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for plates

6.4 Material data sets for plates


In addition to material data sets for soil and interfaces, the material properties and model parameters for plates
are also entered in separate material data sets. Plates are used to model the behaviour of slender walls, plates or
thin shells. Distinction can be made between elastic, elastoplastic and elastoplastic (M-κ) behaviour. A data set
for plates generally represents a certain type of plate material, and can be assigned to the corresponding (group
of) plate elements in the geometry model.

6.4.1 General properties


In the General tabsheet properties with respect to the material behaviour, weight and also damping in a
Dynamics analysis must be specified. Figure 172 (on page 270) shows the General tabsheet of the Plate window.

Figure 172: General tabsheet of the Plate window

6.4.1.1 Material set


The material data set is defined by:

A user may specify any identification title for a data set. It is advisable to use a
Identification meaningful name since the data set will appear in the database tree view by its
identification.

PLAXIS 270 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for plates

There are three available options, describing the material type of a plate. These options
Material type are Elastic, Elastoplastic and Elastoplastic (M-κ). The availability of the parameters
defined at the Mechanical tabsheet depends on the selected material type.

Comments A user may write down comments related to the material data set.

Colour Colour can be used as a distinction tool in the model.

6.4.1.2 Unit weight


Additionally, the unit weight of the plate must be specified:

In a material set for plates a specific weight can be


w [kN/m/m]
specified.

6.4.1.3 Rayleigh damping[ULT]


For dynamic behaviour, the Rayleigh α and Rayleigh β damping parameters can be specified either directly or
through Single Degree of Freedom (SDOF) equivalent parameters.
For more information on Rayleigh damping and how to specify the parameters, see Rayleigh damping [ULT] (on
page 169).

6.4.1.4 Advanced
Prevent punching: In reality vertical loads on walls for example, as a result of vertical components of anchor
forces are sustained by the shaft friction and the tip resistance. A certain amount of resistance is offered by the
soil under the tip, depending on the thickness or the cross section area of the tip.
Slender walls are often modelled as plates. Due to the zero thickness of the plate elements vertical plates (walls)
have no end bearing. The effects of end bearing can still be considered in the calculation when the Prevent
punching option is selected in the material data set. In order to consider end bearing at the bottom of plates, a
zone in the soil volume elements surrounding the bottom of the plate is identified where any kind of soil
plasticity is excluded (elastic zone). The size of this zone is determined as D = 12EI / EA.
It is not the idea that this end bearing represents the true end bearing capacity of the wall. It just prevents
unrealistic vertical movement of penetration of the wall into deeper layers. The value of end bearing cannot be
prescribed by the user, and it is neither theoretically determined by PLAXIS 2D.

Note:
Note that sheet pile walls have very little end bearing, considering that the thickness of the steel is much less
than Deq, when modelling them as plates. Hence, the Prevent punching option shall not be used for sheet pile
walls.

PLAXIS 271 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for plates

6.4.2 Mechanical properties


The mechanical properties required for plates can be grouped into stiffness properties and strength properties
(in case of elastoplastic behaviour).

6.4.2.1 Stiffness properties


For elastic behaviour, several parameters should be specified as material properties. PLAXIS 2D allows for
orthotropic material behaviour in plates where the in-plane stiffness and out-of-plane stiffness are different.
This is most relevant for axisymmetric models when modelling sheet pile profiles (which have a low stiffness in
the out-of-plane direction).

Isotropic If selected, the in-plane stiffness and out-of-plane stiffness will be [-]
equal.

EA1 In-plane elastic axial stiffness. [kN/m]

EA2 Out-of-plane elastic axial stiffness. [kN/m]

Elastic bending stiffness. Note that in the case of Elastoplastic (M-


EI κ) behaviour, EI is automatically determined based on the first [kN/m2/m]
line segment in the M-κ diagram.

ν Poisson's ratio. [-]

From the ratio of EI and EA1 an equivalent thickness for an equivalent plate (d) is automatically calculated from
the equation:
EI
d = 12 EA Eq. [52]
1

For the modelling of plates, PLAXIS 2D uses the Mindlin beam theory as described in Bathe (1982) (on page
552). This means that, in addition to bending, shear deformation is taken into account. The shear stiffness of the
plate is determined from:
5EA 5E (d ⋅ 1m)
Shear stiffness = 12(1 + ν )
= 12(1 + ν )
Eq. [53]

This implies that the shear stiffness is determined from the assumption that the plate has a rectangular cross
section. In the case of modelling a solid wall, this will give the correct shear deformation. However, in the case of
steel profile elements, like sheet-pile walls, the computed shear deformation may be too large. You can check
this by judging the value of deq. For steel profile elements, deq should be at least of the order of a factor 10 times
smaller than the length of the plate to ensure negligible shear deformations.
More information about the behaviour and structural forces in plates can be found in the Material Models
Manual - Structural elements - 2D Plates.
In addition to the above stiffness parameters, a Poisson's ratio, ν, is required. For thin structures with a certain
profile or structures that are relatively flexible in the out-of-plane direction (like sheet-pile walls), it is advisable
to set Poisson's ratio to zero. For real massive structures (like concrete walls) it is more realistic to enter a true
Poisson's ratio of the order of 0.15.

PLAXIS 272 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for plates

Since PLAXIS 2D considers plates (extending in the out-of-plane direction) rather than beams (one-dimensional
structures), the value of Poisson's ratio will influence the bending stiffness of the isotropic plate in plane strain
applications as follows:
Input value of bending stiffness: EI
Observed value of bending stiffness: EI/(1 - ν 2)
The stiffening effect of Poisson's ratio is caused by the stresses in the out-of-plane direction (σzz) and the fact
that strains are prevented in this direction. Note that the Poisson's ration (ν) is assumed to be zero in anisotropic
case.

6.4.2.2 Strength properties


Elastoplastic
Strength parameters are required in the case of elastoplastic behaviour:

Mp Maximum bending moment. [kNm/m]

Np,1 The maximum axial force in 1-direction. [kN/m]

Np,2 The maximum axial force in 2-direction (anisotropic behaviour). [kN/m]

Plasticity may be taken into account by specifying a maximum bending moment, Mp. The maximum bending
moment is given in units of force times length per unit width. In addition to the maximum bending moment, the
axial force is limited to Np. The maximum axial force, Np, is specified in units of force per unit width. When the
combination of a bending moment and an axial force occur in a plate, then the actual bending moment or axial
force at which plasticity occurs is lower than respectively Mp or Np. More information is given in the Material
Models Manual - Structural elements - 2D Plates.
The relationship between Mp and Np is visualised in Figure 173 (on page 274). The diamond shape represents
the ultimate combination of forces for which plasticity will occur. Force combinations inside the diamond will
result in elastic deformations only. The Scientific Manual describes in more detail how PLAXIS 2D deals with
plasticity in plates.
Bending moments and axial forces are calculated at the stress points of the plates elements (Figure 62 (on page
103)) . If Mp or Np is exceeded, stresses are redistributed according to the theory of plasticity, so that the maxima
are complied with. This will result in irreversible deformations. Output of bending moments and axial forces is
given in the nodes, which requires extrapolation of the values at the stress points. Due to the position of the
stress points in a beam element, it is possible that the nodal values of the bending moment may slightly exceed
Mp. If the Isotropic option is checked the input is limited to Np,1 where as Np,1 = Np,2.

PLAXIS 273 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for plates

Np

Mp Mp

Np

Figure 173: Combinations of maximum bending moment and axial force

Elastoplastic (M-κ)
In the case the Material type is set to Elastoplastic (M-κ), the bending behaviour of the plate is governed by a
user-defined moment-curvature (M-κ) diagram. Input values of the κ diagram can be specified in the table that
appears at the right hand panel when the M-κ diagram input field is selected. Here, a name can be given to the M-
κ diagram to be defined. Adding rows to the table can be done by clicking the '+' sign above the table. Only
positive values can be specified for curvature κ (in the unit of 1/length) and moment M (in the unit of Force
times length per unit of width in the out-of-plane direction), since the M-κ diagram is assumed to govern the
plate's behaviour in both positive and negative loading directions. A graphical representation of the M-κ diagram
is presented below the table (only positive values).
The first row in the M-κ diagram determines the plate's elastic bending stiffness EI:
EI = M (1) / κ (1) Eq. [54]
This EI value is included in the set of stiffness properties. Additional rows in the M-κ table define the plate's non-
linear elastoplastic behaviour. The last M-value in the table is assumed to be the maximum bending moment, Mp,
at which the plate fails. Elastic bending strains are calculated from the aforementioned EI value, whereas
additional bending (following the defined M-κ diagram) is assumed to be plastic, as long as the loading is in the
same direction. Upon unloading from an elastoplastic state, the behaviour is initially elastic, but after significant
unloading it may again generate plastic strains.
In the Elastoplastic (M-κ) plates, the axial force N is considered elastic and independent from the bending
moment M. The axial force influences the stresses and displacements of the element but without affecting the
bending moment. The M-κ diagram is assumed to be the only one governing the plate's flexural behaviour.

M1

Mp

E1

kappa

Figure 174: M-κ diagram

PLAXIS 274 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for geogrids

It is possible to change the material data set of a plate in the framework of staged construction . However, it is
very important that the ratio of EI/EA is not changed, since this will introduce an out-of-balance force (see Plates
(on page 101)).

6.4.3 Thermal properties [ADV]


For thermal behaviour, additional constants are specified in the Thermal tabsheet.

The specific heat capacity of the solid material. Describes the amount of energy
c [kJ/t/K]
(heat) that can be stored in the solid material per unit of mass.

The thermal conductivity of the solid material. Describes the rate of energy
λ [kW/m/K]
(heat) that can be transported in the solid material.

ρ The density of the solid material. Describes the density of the solid particles. [t/m3]

The thermal expansion coefficient. Describes how much the material expands (or
α elongates) when the temperature increases. In other words, it is the (change of) [1/K]
strain per unit of temperature.

Aeff,T The effective area for thermal flux. [m2]

6.5 Material data sets for geogrids


In addition to material data sets for soil and interfaces, the material properties and model parameters for
geogrids are also entered in separate material data sets. Geogrids are flexible elastic or elastoplastic elements
that represent a grid or sheet of fabric. Geogrids can only sustain tensile forces, but not compressive forces. A
data set for geogrids generally represents a certain type of geogrid material, and can be assigned to the
corresponding (group of) geogrid elements in the geometry model.

6.5.1 General properties


Several data sets may be created to distinguish between different types of geogrids or geosynthetics. Figure 175
(on page 276) shows the General tabsheet of the Geogrid window. The general properties of a geogrid are:

A user may specify any identification title for a data set. It is advisable to use a
Identification meaningful name since the data set will appear in the database tree view by its
identification.

Comments A user may write down comments related to the material data set.

Colour Colour can be used as a distinction tool in the model.

PLAXIS 275 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for geogrids

There are four available options, describing the material type of a plate. These options
are Elastic, Elastoplastic, Elastoplastic (N-ε) and Visco-elastic (time-dependent). The
Material type
availability of the parameters defined at the Mechanical tabsheet depends on the
selected material type.

Figure 175: Geogrid window

6.5.2 Mechanical properties


The properties required for geogrids can be grouped into stiffness properties and strength properties in case of
elastoplastic behaviour.

6.5.2.1 Isotropic
Different stiffness and strength in-plane and out-of-plane may be considered. If this is not the case, the Isotropic
option may be selected to ensure that the stiffness in both directions is equal.

6.5.2.2 Stiffness properties


Elastic
For elastic behaviour, the axial stiffness EA should be specified. PLAXIS 2D allows for orthotropic material
behaviour in geogrids, which is defined by the following parameters:

EA1 The axis axial elastic stiffness in 1-direction (in-plane). [kN/m]

The axial elastic stiffness in 2-direction (out-of-plane, anisotropic


EA2 [kN/m]
behaviour).

PLAXIS 276 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for geogrids

The axial stiffness EA is usually provided by the geogrid manufacturer and can be determined from diagrams in
which the elongation of the geogrid is plotted against the applied force in a longitudinal direction. The axial
stiffness is the ratio of the axial force F per unit width and the axial strain (Δl/l where Δl is the elongation and l is
the original length):
F
EA = Δl / l
Eq. [55]

If the Isotropic option is checked the input is limited to EA1 where as EA2 =EA1.
Visco-elastic (time-dependent)
The parameters are based on a visco-elastic perfectly-plastic Kelvin-Voigt model in each direction, which allows
for time-dependent behaviour.
F

u
Figure 176: Force versus displacement showing EAshort and EAlong

Considering one Kelvin-Voigt element (see Figure 177 (on page 277)),

Figure 177: Kelvin-Voigt single element representation

E A0 = E Ashort
1 1 1
E A0
+ E A1
= E Along

1 Eq. [56]
E A1 =
( 1
E Along

1
E A0 )
η = E A1 × Retardation time

where η is viscous damping, EA0 and EA1 are internal stiffness used in the geogrid material model.

PLAXIS 277 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for geogrids

Figure 178: Displacement versus time in a Creep test

Retardation time can be obtained from a creep test in which a force F is applied, using the following equations
(see Figure 178 (on page 278)):
η1
tretardation = E1

F
E Ashort = ushort

1
E Along = E Ashort
Eq. [57]

F
ushort = E Ashort

F
ulong = E Along

The short-term and long-term stiffnesses can be obtained from the same creep test by measuring the
corresponding displacement u at the short and long time, respectively.
Parameters which are required for time dependent visco-elasticity are (Figure 179 (on page 279)):

EA1,short Elastic stiffness during initial (instantaneous) strain increment in 1- [kN/m]


direction (in-plane).

EA2,short Elastic stiffness during initial (instantaneous) strain increment in 2- [kN/m]


direction (out-of-plane, anisotropic behaviour).

EA1,long Elastic stiffness during (infinitely) long strain increment in 1-direction [kN/m]
(in-plane).

EA2,long Elastic stiffness during (infinitely) long strain increment in 1-direction [kN/m]
(out-of-plane, anisotropic behaviour).

Retardation time The time where a linear extrapolation of the initial creep rate intersects [day]
the long-term displacement line.

PLAXIS 278 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for geogrids

Figure 179: Geogrid window

6.5.2.3 Strength properties


Elastoplastic
Strength parameters are required in case of elastoplasticity:

Np,1 The maximum force in 1-direction (in-plane). [kN/m]

The maximum force in 2-direction (out-of-plane, anisotropic


Np,2 [kN/m]
behaviour).

The maximum axial tension force Np is specified in units of force per unit width. If Np is exceeded, stresses are
redistributed according to the theory of plasticity, so that the maxima are complied with. This will result in
irreversible deformations. Output of axial forces is given in the nodes, which requires extrapolation of the values
at the stress points. Due to the position of the stress points in a geogrid element, it is possible that the nodal
values of the axial force may slightly exceed Np.
If the Isotropic option is checked the input is limited to Np,1 where as Np,2 =Np,1.
Elastoplastic (N-ε)
A non-linear N-ε diagram may be specified in case of elastoplasticity (N-ε):

N1-ε1 diagram The N-ε diagram in 1-direction (in-plane).

N2-ε2 diagram The N-ε diagram in 2-direction (out-of-plane, anisotropic behaviour).

The user can add both N-ε diagrams, by specifying values as a table or importing a text file containing values
(Figure 180 (on page 280)). The axial tension force is specified in units of force per unit width. The first part of
the N-ε diagram is used to calculate the EA value.
If the Isotropic option is checked the input is limited to N1 -eps1 where as N2-eps2 =N1-eps1.

PLAXIS 279 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for embedded beams

Figure 180: Geogrid window

6.5.3 Thermal properties [ADV]


For thermal behaviour, additional constants are specified in the Thermal tabsheet.

The specific heat capacity of the solid material. Describes the


c amount of energy (heat) that can be stored in the solid material per [kJ/t/K]
unit of mass.

The thermal conductivity of the solid material. Describes the rate of


λ [kW/m/K]
energy (heat) that can be transported in the solid material.

The density of the solid material. Describes the density of the solid
ρ [t/m3]
particles.

The thermal expansion coefficient. Describes how much the


α material expands (or elongates) when the temperature increases. In [1/K]
other words, it is the (change of) strain per unit of temperature.

Aeff,T The area in consideration for the thermal expansion. [m2]

Note: The creep properties of geogrids are very sensitive to the temperature. The visco-elastic perfectly-plastic
model adopted for geogrids allows time-dependent behaviour but without temperature-dependency. Hence, the
thermal expansion coefficient does not take the temperature effect on creep into account.

PLAXIS 280 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for embedded beams

6.6 Material data sets for embedded beams


Properties and model parameters for embedded beam are entered in separate material data sets. Embedded
beams can be used to model different types of slender structures that interact with the surrounding soil/rocks
such as piles, ground anchors or rock bolts. A data set for embedded beams generally represents a certain type
of embedded beam, including the pile/rock bolt material and geometric properties, the interaction properties
with the surrounding soil or rock (bearing capacity) as well as the out-of-plane spacing of the piles/rock bolts.
Note that the embedded beam material data set does not contain so-called 'p-y curves', nor equivalent spring
constants. In fact, the stiffness response of an embedded beam subjected to loading is the result of the specified
embedded beam length, equivalent radius, spacing, stiffness, bearing capacity, the stiffness of the interface as
well as the stiffness of the surrounding soil.

Note:
In contrast to what is common in the Finite Element Method, the bearing capacity of an embedded beam is
considered to be an input parameter rather than the result of the finite element calculation. The user should
realise the importance of this input parameter. Preferably, the input value of this parameter should be based on
representative pile load test or pull out test data. Moreover, it is advised to perform a calibration in which the
behaviour of the embedded beam is compared with the behaviour as measured from the test. When embedded
beams are used in a row, the group action must be taken into account when defining their bearing capacity.

Note: Embedded beams can only be used in plane strain models and are not available in axisymmetry.

6.6.1 General properties


Several data sets may be created to distinguish between different types of embedded beams. Figure 181 (on
page 282) shows the Embedded beam window.

PLAXIS 281 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for embedded beams

Figure 181: Embedded beam window

6.6.1.1 Material set


The material data set is defined by:

A user may specify any identification title for a data set. It is advisable to use a
Identification meaningful name since the data set will appear in the database tree view by its
identification.

There are three available options, describing the material type for embedded beams.
Material type These options are Elastic , Elastoplastic and Elastoplastic (M-κ). The availability of the
parameters defined on the Mechanical tabsheet depends on the selected material type.

Colour Colour can be used as a distinction tool in the model.

Comments A user may write down comments related to the material data set.

6.6.1.2 Unit weight


Additionally, the unit weight of the embedded beam must be specified:

The unit weight is the unit weight of the material from which the
γ [kN/m3]
embedded beam is composed.

PLAXIS 282 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for embedded beams

6.6.1.3 Rayleigh damping [ULT]


For dynamic behaviour, the Rayleigh α and Rayleigh β damping parameters can be specified either directly or
through Single Degree of Freedom (SDOF) equivalent parameters.
For more information on Rayleigh damping and how to specify the parameters, see Rayleigh damping [ULT] (on
page 169).

6.6.2 Mechanical properties


The material properties are defined for a single beam, but the use of PLAXIS 2D implies that a row of piles or
rock bolts in the out-of-plane direction is considered where properties are specified per unit width in out-of-
plane direction.
Figure 182 (on page 284) shows the Mechanical tabsheet for an embedded beam in PLAXIS 2D.

PLAXIS 283 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for embedded beams

Figure 182: Mechanical tabsheet of the Embedded beam window

Note:
• In a finite element model, embedded beams are superimposed on a continuum and therefore ' overlap' the
soil. Especially for Solid structures, to calculate accurately the total weight of soil and structures in the model,
the unit weight of the soil should be subtracted from the unit weight of the embedded beam material.
• Please note that when reducing the unit weight, the axial forces in the embedded beams may not be realistic.

6.6.2.1 Cross section properties


An embedded beam requires several geometric parameters used to calculate additional properties. First of all,
the out-of plane spacing Lspacing in [m] must be defined in order to calculate all parameters per meter width out-
of-plane.

PLAXIS 284 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for embedded beams

Cross section type Either a Predefined or a user-defined type can be selected.

Predefined cross A list of predefined types (Solid circular beam, Circular tube, Solid square beam). After
section type entering the required data, the parameters A, I3, I2, W3 and W2 are filled automatically ).

The embedded beam diameter is to be defined for Solid circular beam and Circular tube
predefined embedded beam types. The embedded beam diameter determines the size of
Diameter the elastic zone in the soil under the embedded beam in which plastic soil behaviour is
excluded. It also influences the default values of the interface stiffness factors (Interface
stiffness factor (on page 289)).

The embedded beam width is to be defined for a Solid square beam predefined beam
type. The embedded beam width is recalculated into an equivalent diameter,
Width Deq = 12EI / EA. This diameter determines the size of the elastic zone in the soil
under the embedded beam in which plastic soil behaviour is excluded. It also influences
the default values of the interface stiffness factors.

Thickness The wall thickness needs to be defined for a Circular tube predefined beam type.

Alternatively, a user-defined type may be defined by means of the embedded beam cross section area, A, and its
respective moment of inertia I:

The cross section area is the actual area (in the unit of length
squared) perpendicular to the embedded beam axis where
A embedded beam material is present. For embedded beams that [m2]
have a certain profile (such as steel beams), the cross section area
can be found in tables that are provided by the manufacturer.

Moment of inertia against bending around the embedded beam


I [m4]
axis.

1
l= 64
πD 4 Solid circular embedded beam

l=
1
4
( D2 )4 − ( D2 − t )4 Circular tube Eq. [58]
1
l= 12
h4 Solid square embedded beam

where D is the embedded beam diameter, t is the wall thickness and h is the embedded beam width.}

6.6.2.2 Stiffness
An embedded beam has only a single stiffness parameter:

E Young's modulus. [kN/m2]

6.6.2.3 Strength
Elastoplastic

PLAXIS 285 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for embedded beams

For elastoplastic behaviour of an embedded beam, the strength is defined by the plastic limits for bending
moment and axial force:

Mp Maximum (plastic) bending moment [kNm]

Np Maximum (plastic) axial force [kN]

Elastoplastic (M-κ)
In the case the Material type is set to Elastoplastic (M-κ), the bending behaviour of the plate is governed by a
user-defined moment-curvature (M-κ) diagram. Input values of the κ diagram can be specified in the table that
appears at the right hand panel when the M-κ diagram input field is selected.
For more information on elastoplastic behaviour see Strength properties (on page 273)

6.6.3 Interaction properties: bearing capacity


The interaction between the pile or rock bolt (embedded beam element) and the surrounding soil or rock (soil
volume element) is modelled by means of a special interface element. An elastoplastic model is used to describe
the behaviour of the interface. The elastic behaviour of the interface should account for the difference in pile/
rock bolt displacements and average soil/rock displacements in the out-of-plane direction. This depends on the
out-of-plane spacing in relation to the diameter. Regarding the plastic behaviour, distinction is made in the
material data set between the Skin resistance (in the unit of force per unit embedded beam length) and the Base
resistance (in the unit of force). In a plane strain analysis, these values are automatically recalculated per unit of
width in the out-of-plane direction.For the skin resistance, as well as the base resistance, a failure criterion is
used to distinguish between elastic interface behaviour and plastic interface behaviour. For elastic behaviour
relatively small displacement differences occur within the interface (i.e. between the pile/rock bolt and the
average soil/rock displacement), and for plastic behaviour permanent slip may occur.
For the interface, to remain elastic, the shear force ts at a particular point is given by:
| ts | < T max Eq. [59]

where Tmax is the equivalent local skin resistance at that point.


For plastic behaviour the shear force ts is given by:
| ts | = T max Eq. [60]

The input for the shaft resistance is defined by means of the axial skin resistance Fskin and the base resistance
Fmax. Using this approach the total pile bearing capacity, Npile, is given by:
N pile = F max + F skin Eq. [61]

The skin resistance Fskin can be defined in three ways: Linear, Multi-linear and Layer dependent. The base
resistance Fmax can be entered directly (in the unit of force) in the embedded beam material data set window.

Note:
The base resistance is only mobilised when the pile body moves in the direction of the base (example: with a
load on top).

For rock bolts, the bearing capacity may be defined in a similar way, although rock bolts generally do not have an
end bearing (Fmax = 0):

PLAXIS 286 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for embedded beams

N rockbolt = F skin Eq. [62]

In order to ensure that a realistic bearing capacity as specified can actually be reached, a zone in the soil volume
elements surrounding the beam is identified where any kind of soil plasticity is excluded (elastic zone). The size
of this zone is determined by the embedded beam's diameter D or equivalent diameter Deq. The elastic zone
makes the embedded beam almost behave like a volume element. Installation effects are not taken into account
and the embedded beam-soil interaction is modelled at the centre rather than at the circumference.
The bearing capacity is automatically divided by the spacing in order to obtain the equivalent bearing capacity
per unit of width in the out-of-plane direction.
In addition to displacement differences and shear forces in axial direction along the embedded beam element,
the beam can undergo transverse forces, t⊥ , due to lateral displacements. The lateral displacements can be
induced by a transverse force applied at the top of the pile or as a consequence of the transverse distributed load
induced by the lateral displacement field of the surrounding soil. In the first case, the overall behaviour may not
show fully realistic results. In the second case, although embedded beam elements are not meant to be used as
laterally loaded piles, they show reasonable results and the overall behaviour is quite realistic. For more
information about the modelling techniques for embedded structures, see the Appendix of the Material Models
Manual.
The user can limit the transverse forces using the Lateral resistance option. This can be done in three ways:
Unlimited, Linear and Multi-linear lateral resistance.
More details about the way the shear and transverse forces are calculated on the basis of displacement
differences between the embedded beam element and the surrounding soil element are described in the
Material Models Manual.

6.6.3.1 Axial skin resistance


Linear
The Linear option is the easiest way to enter the skin resistance and lateral resistance profile. The input is
defined by means of the skin resistance and lateral resistance at the beginning of the embedded beam,
Tskin,start,max and Tlat,start,max (in force per unit beam length), and the skin resistance and lateral resistance at the
end of the embedded beam, Tskin,end,max and Tlat,end,max (in force per unit beam length). This way of defining the
skin resistance and lateral resistance is mostly applicable to piles in a homogeneous soil layer. Using this
approach the total pile bearing capacity, Npile, is given by:
1
N pile = F max + 2
L pile (T skin,start ,max + T skin,end ,max ) Eq. [63]

where Lpile is the pile length. Tskin,start,max and Tskin,end,max are measured at the pile top and the bottom of the pile
respectively.
For rock bolts, the bearing capacity may be defined as follows:
1
N rockbolt = 2
L rockbolt (T skin,start ,max + T skin,end ,max ) Eq. [64]

where Lrockbolt is the rock bolt length. Tskin,start,max and Tskin,end,max are measured at the first point of the geometry
line and the second one respectively.
Multi-linear
The Multi-linear option can be used to take into account inhomogeneous or multiple soil layers with different
properties and, as a result, different resistances. The skin resistance, Tmax, is defined in a table at different

PLAXIS 287 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for embedded beams

positions along the pile, L, where L is measured from the pile top (L=0) to the bottom of the pile (L= Lpile). Using
this approach the total pile bearing capacity, Npile, is given by:
n−1 1
N pile = F max + ∑ 2 ( L i+1 − L i )(T i + T i+1 ) Eq. [65]
i=1

where i is the row number in the table and n is the total number of rows.
For rock bolts, the bearing capacity is defined as follows:
n−1 1
N rockbolt = ∑ 2 ( L i+1 − L i )(T i + T i+1 ) Eq. [66]
i=1

where L is measured from the first point of the corresponding geometry line (L=0) to the second one (L=
Lrockbolt).
Layer dependent
The Layer dependent option can be used to relate the local skin resistance to the strength properties (cohesion c
and friction angle φ) and the interface strength reduction factor, Rinter, as defined in the material data set of the
corresponding soil or rock layers ( Interfaces tabsheet (on page 247)) in which the pile or rock bolt is located:
τi = ci + σ ′ n tan φi
ci = Rintercsoil Eq. [67]
tan φi = Rintertan φsoil

where τi is the local shear stress resistance of the interface, φi and ci are the friction angle and the cohesion of
the interface, φsoil and csoil are the friction angle and cohesion of the correspondent soil layer, Rinter is the
strength reduction factor associated to the soil layer, σ'n = (σ'2 + σ'3 ) / 2 is the normal stress. Using this approach
the pile/rock bolt bearing capacity is based on the stress state in the soil, and thus unknown at the start of a
calculation. The special interface in the embedded beam behaves similar as an interface along a wall, except that
it is a line interface rather than a sheet.

Note:
In case of advanced soil material models, the calculation of φi and ci is as described in Interfaces tabsheet (on
page 247). For user-defined Soil Models, φi and ci are based on direct input specified by the user.

The skin resistance, Ti, as a force per unit of depth, is defined as:
T i = 2πReq τi Eq. [68]

where Req is the embedded beam's equivalent radius.


To avoid that the skin resistance could increase to undesired high values, an overall maximum resistance Tmax
(constant value along the pile/rock bolt in force per unit pile/rock bolt length) can be specified in the embedded
beam material data set, which acts as an overall cut-off value.

Note:
The embedded beam-soil interaction parameters in the embedded beam material data set involve only the
bearing capacity (skin resistance and base resistance). Note that the material data set does NOT include the
stiffness response of the pile in the soil (or p-y curve). The stiffness response is the result of the pile length,
equivalent radius, stiffness and bearing capacity as well as the stiffness of the soil layers in which the beam is
located.

PLAXIS 288 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for embedded beams

6.6.3.2 Lateral resistance


Unlimited
The value of the lateral resistance in this case is unlimited.
Linear
Similar to the linear axial skin resistance the start and end value, Tlat,start,max and Tlat,end,max must defined.
Multi-linear
Again similar to the multi-linear axial skin resistance a table of lateral skin resistance as a function of the
distance along the beam can be specified.

Figure 183: Available options for Lateral resistance

6.6.3.3 Base resistance


The base resistance Fmax is the maximum foot force of the embedded beam, hence the end-bearing capacity. Note
that the base of the embedded beam can only sustain a compressive force and no tensile force.

6.6.4 Interface stiffness factor


The interface stiffnesses are related to the shear stiffness of the surrounding soil (Gsoil) according to:

PLAXIS 289 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for embedded beams

Gsoil
Rs = ISF RS L spacing

Gsoil
R N = ISF RN L spacing
Eq. [69]

Gsoil Req
K F = ISF KF L spacing

Figure 184: Modelling of soil-pile interaction

The interface stiffness factors to be defined are:


• Axial skin stiffness factor, ISFRS
• Lateral stiffness factor, ISFRN
• Pile base stiffness factor, ISFKF
where the default values of the interface stiffness are calculated according to:

ISF RS = 25 ( )
L spacing −0.75
D

= 25( )
L spacing −0.75
ISF RN Eq. [70]
D

= 25( )
L spacing −0.75
ISF KF D

where D is diameter in the case of a circular pile or width in the case of a square pile or equivalent width in the
case the case of a user-defined pile.
In the case of an embedded beam as a pile a realistic pile bearing capacity as specified can actually be reached, a
zone in the soil volume elements surrounding the bottom of the pile is identified where any kind of soil plasticity
is excluded (elastic zone). The size of this zone is determined by the embedded beam's diameter Deq or
equivalent radius Req (= Deq/2) (Figure 185 (on page 291))
where
12EI
Deq = EA
Eq. [71]

PLAXIS 290 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for cable bolts

Figure 185: Plastic zone surrounding the bottom of the pile (after Sluis, 2012)

In addition to displacement differences and shear forces in axial direction along the pile, the pile can undergo
transverse forces, t⊥ , due to lateral displacements. These transverse forces are not limited in the special
interface element that connects the pile with the soil, but, in general, they are limited due to failure conditions in
the surrounding soil itself. However, embedded beams are not meant to be used as laterally loaded piles and will
therefore not show accurate failure loads when subjected to transverse forces.
Note that the default values of the interface stiffness factors are valid for bored piles which are loaded statically
in the axial direction and behaviour of the surrounding soil is modelled using the Hardening Soil model with
small-strain stiffness. The phreatic level is assumed to be located at the ground surface. These values should be
modified if the conditions in the model are different from the ones assumed to derive the default values.

6.7 Material data sets for cable bolts


As mentioned in the chapter Modelling loads and structures Structures mode - Cable bolts (on page 109), the
material properties for this element can be entered in the material data set window under 2 tab sheets: General
and Mechanical.

6.7.1 General properties

Several data sets may be created to distinguish different cable bolts. Figure 186 (on page 292) displays the
General properties tab for Cable bolts.

PLAXIS 291 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for cable bolts

Figure 186: General material properties of cable bolts

Users may specify any identification title for a cable bolt data set. It is advised to use a
Identification meaningful name since the data set will appear in the database tree view by its
identification.

Material Model Users can specify the elastic behaviour of the cable bolt as "Elastic" or as "Elastoplastic".

Colour can be used as a distinction tool in the model. The default color for a cable bolt is
Color
purple.

Comments A user may write down comments related to the material data set.

Note: When a elastic behaviour is selected for the cable bolts different parameters become available in the
Mechanical properties (on page 292) tab.

6.7.2 Mechanical properties

The required parameters for the strength and stiffness of cable bolts are defined in the Mechanical tabsheet of
the material set as shown in Figure 187 (on page 293).

PLAXIS 292 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for cable bolts

Figure 187: Mechanical material properties of cable bolts

6.7.2.1 Cross Section properties


A cable bolt requires several geometric parameters used to calculate additional properties.

Refers to the out-of plane spacing of cable blots, it must be defined in order to
Lspacing m
calculate all parameters per unit of length in the out-of-plane.

PLAXIS 293 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for cable bolts

Cross
Either a Predefined or a user-defined type can be selected. -
section type

A list of predefined types:


• Solid circular beam:
Requires to fill the Diameter of the cable and the Area (A) is calculated
automatically.
Predefined
cross section The cable bolt diameter determines the size of the elastic zone in the rock m
type around the cable bolt in which plastic rock behaviour is excluded. It also
influences the default values of the interface stiffness factors.
• Circular tube:
Requires to fill the Diameter and the Thickness of the tube to calculate the Area
(A) of the cable bolt.

The cross section area is the actual area (in the unit of length squared)
A m2
perpendicular to the cable bolt axis where the cable bolt material is present.

6.7.2.2 Cable Stiffness


E Axial stiffness of the cable bolt kN/m2

6.7.2.3 Cable Strength [only for elastoplastic cable bolts]


Np, comp Compression limit of the cable kN

Np, tens Tension limit of the cable kN

6.7.2.4 Bond Stiffness


Ks Stress dependent shear stiffness of the bond kN/m2

6.7.2.5 Bond Strength


The bond strength capacity is calculated by:
T s,bond,max = cbond + σn tan(ϕbond) * perimeter Eq. [72]
where
σn = Confining stress normal to the cable axis

other strength parameters are defined as follows:

PLAXIS 294 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for cable bolts

When activated no bond failure (plasticity) takes place. Also the strength
Fully bonded distribution is set as Uniform by default with high value stiffness and strength -
parameters.

If the cable is NOT fully bonded the following properties are displayed:
a) Strength distribution set as Uniform:

Cohesive
strength Bond cohesive strength. kN/m
cbond

φbond Bond friction angle to account for the bond confining pressure dependency. °

• Predefined:
The failure perimeter is calculated by default based on cable diameter,
assuming that failure takes place at the cable circumference.
• User-defined:
The users need to assign a value for the failure perimeter.
To define the failure parameter take into account that bond failure takes place
at the interface between the reinforcement element and medium. When
frictional resistance and confining pressure effects are considered, the failure
perimeter is needed to calculate the surface of that interface. Two approaches
are possible to define the failure perimeter:
Failure
surface • Non-grouted bolt: m
perimeter For non-grouted rock bolt, failure perimeter can be taken as the perimeter
of the bolt cross section.
• Grouted bolt and cable:

For grouted bolt and cable, PLAXIS assumes, by default, that failure takes
place at grout-bolt interface, so the predefined perimeter is also the
perimeter of the bolt cross section.
However, if other bond failure mechanism is assumed, user-defined
perimeter option can be used, so that users can put their own failure
perimeter value. For instance, if failure at rock-grout interface is assumed,
the failure perimeter can be taken as the perimeter of the drilling hole, used
to install the bolt/cable.

b) Strength distribution set as Multilinear:


Instead of having a uniform value, shear strength parameters (cohesive strength, friction angle) can vary along
the cable length, starting from the first point of the cable geometry line (Cable bolt properties (on page 111) ).

Multi-linear
shear Predefined as 2 rows. kN
strength

PLAXIS 295 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for anchors

Failure
surface • Predefined m
perimeter • User-defined

Note: For the safety analysis a Phi/C reduction all strength parameters as part of the cable bolt will be reduced,
including the rock bolt interaction parameters (Np, comp, Np, tens, Cohesive strength, φbond ). But it is possible to
exclude those parameters from the reduction domain.

6.8 Material data sets for anchors


In addition to material data sets for soil and interfaces, the material properties and model parameters for
anchors are also entered in separate material data sets. A material data set for anchors may contain the
properties of node-to-node anchors as well as fixed-end anchors. In both cases the anchor is just a spring
element. A data set for anchors generally represents a certain type of anchor material, and can be assigned to the
corresponding (group of) anchor elements in the geometry model.
Anchors can be prestressed in a Staged construction calculation. In such a calculation the prestress force for a
certain calculation phase can directly be given in the Object explorer. The prestress force is not considered to be
a material property and is therefore not included in an anchor data set.

PLAXIS 296 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for anchors

6.8.1 General properties

Figure 188: Anchor window

Several data sets may be created to distinguish between different types of anchors. Figure 188 (on page 297)
shows the General tabsheet of the Anchor window. The material data set is defined by:

A user may specify any identification title for a data set. It is advisable to use a
Identification meaningful name since the data set will appear in the database tree view by its
identification.

There are three available options, describing the material type of an anchor. These
options are Elastic, Elastoplastic and Elastoplastic with residual strength. The availability
Material type
of the parameters defined on the Mechanical tabsheet depends on the selected material
type.

Colour Colour can be used as a distinction tool in the model.

Comments A user may write down comments related to the material data set.

6.8.2 Mechanical properties


The properties required for anchors can be grouped into stiffness properties and strength properties in case of
elastoplastic behaviour.

PLAXIS 297 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for anchors

6.8.2.1 Stiffness
To calculate an equivalent stiffness per unit width, the out-of-plane spacing, Ls, must be entered.

Note: For axisymmetric problems Ls is considered in radians and not in unit width.

An anchor requires only one stiffness parameter:

EA Axial stiffness of a single anchor. [kN]

6.8.2.2 Strength
Elastoplastic
If the material type is selected as Elastoplastic, two maximum anchor forces can be entered:

Fmax,tens Maximum tension force of a single anchor [kN]

Fmax,comp Maximum compression force of a single anchor [kN]

The Force-displacement diagram displaying the elastoplastic behaviour of the anchors is given in Figure 189 (on
page 298).

Figure 189: The force-displacement diagram displaying the elastoplastic behaviour of anchors

In the same way as the stiffness, the maximum anchor forces are divided by the out-of-plane spacing in order to
obtain the proper maximum force in a plane strain analysis.
Elastoplastic with residual strength

PLAXIS 298 PLAXIS 2D-Reference Manual


Material properties and material database
Material data sets for anchors

The Elastoplastic with residual strength option can be used to model anchor failure or softening behaviour (e.g.
buckling of struts). When this option is selected two residual anchor forces can be specified:

Fmax,tens Maximum tension force of a single anchor [kN]

Fmax,comp Maximum compression force of a single anchor [kN]

Fresidual,tens Residual tension force of a single anchor [kN]

Fresidual,comp Residual compression force of a single anchor [kN]

The Force-displacement diagram displaying the elastoplastic behaviour with residual strength of the anchors is
given in Figure 189 (on page 298).

Figure 190: The force-displacement diagram displaying the elastoplastic behaviour with residual strength of the
anchor

If, during a calculation, the maximum anchor force is reached, the maximum force will immediately reduce to the
residual force. From that point on the anchor force will not exceed the residual force anymore. Even if the anchor
force would intermediately reduce to lower values, the defined residual force will be its maximum limit.

Note: If the anchor has failed (in tension, compression or both) the residual force will be valid in the following
calculation phases where the anchor is active. If the anchor is deactivated in a phase and reactivated in the next
phase, the maximum anchor force will be restored, assuming that the anchor has been replaced by a new anchor.

Note: Using reduced residual strength is not recommended in Safety calculations.

6.8.3 Thermal properties [ADV]


For thermal behaviour, additional constants are specified in the Thermal tabsheet.

PLAXIS 299 PLAXIS 2D-Reference Manual


Material properties and material database
Assigning material data sets to geometry components

The specific heat capacity of the solid material. Describes the amount of
c [kJ/t/K]
energy (heat) that can be stored in the solid material per unit of mass.

The thermal conductivity of the solid material. Describes the rate of energy
λ [kW/m/K]
(heat) that can be transported in the solid material.

ρ The density of the solid material. [t/m3]

The thermal expansion coefficient. Describes how much the material expands
α (or elongates) when the temperature increases. In other words, it is the [1/K]
(change of) strain per unit of temperature.

Aeff,T The cross sectional area of the anchor considered for thermal flux. [m2]

6.9 Assigning material data sets to geometry components


After creating material data sets, the data sets must be assigned to the corresponding geometry components
(soil layers and structures). This can be done in different ways, which are explained below. The methods
described below are primarily meant to assign properties to the initial geometry. For details on the change of
properties during calculations in the framework of staged construction (Reassigning material data sets (on page
374)).

6.9.1 Soil layers


Regarding soil data, material data sets can be assigned to individual soil layers in the boreholes. Therefore a
borehole should be double-clicked to open the corresponding Modify soil layers window. In the Modify soil
layers click the plus button next to the soil layer. The Material dataset window pops up, where the properties
of the new material can be defined.
Materials button at the lower right hand side of the window should be clicked to open the material database. To
assign a data set to a particular soil layer, select the desired data set from the material database tree view (click
on the data set and hold the left hand mouse button down), drag it to the soil column in the borehole window
(hold the mouse button down while moving) and drop it on the desired layer (release the mouse button). The
layer should now show the corresponding material data set colour. The drag and drop procedure should be
repeated until all layers have their appropriate data set. Note that material sets cannot be dragged directly from
the global database tree view and must be copied to the project database first.
When multiple boreholes are used it should be noted that assigning a data set to a layer in one particular
borehole will also influence the other boreholes, since all layers appear in all boreholes, except for layers with a
zero thickness.

6.9.2 Structures
Regarding structures (fixed-end anchors, node-to-node anchors, embedded beam rows, plates, geogrids and
interfaces), there are three different methods of assigning material data sets.

PLAXIS 300 PLAXIS 2D-Reference Manual

You might also like