• 검색 결과가 없습니다.

Simplified method

문서에서 S a f e t y R e p o r t s S e r i e s N o. 8 7 (페이지 101-107)

4. STRUCTURAL RESPONSE ANALYSIS

4.3. Structural response analysis for explosion loads

4.3.3. Simplified method

4.3.3.1. Simplified versus complex models

In many cases, structural components subject to blast load can be modelled as an equivalent SDOF mass–spring system, with a non-linear spring (Fig. 29).

In creating an equivalent SDOF structure, it is to be realized that the real structure is a multi-degree of freedom system in which every mass particle has its own equation of motion. Thus, to simplify the situation, it is necessary to make kinematical assumptions about the response and, in particular, on how to characterize global deformation in terms of a single point displacement.

Complex models usually involve non-linear finite element computations, in which non-linear geometric and material behaviour is considered and failure criteria implemented. Non-linear finite element computations may start from the load function definition, that is, predefined load functions that vary over space and time, and that are applied to the SSC of interest; or they may start from the explosion or blast itself, that is, from the position, amount, geometry and nature of the explosive material. In the latter case, the blast wave generation, propagation up to the structure and the interaction with it also need to be modelled. This normally requires the combination of Lagrangian and Eulerian processors in the computer code. In many commercial finite element codes, the pressure waves generated by the release of chemical energy in an explosion can be modelled by using, for example, the Jones–Wilkins–Lee equation of state. Special care needs to be paid to verify the calculation method in simulating pressure loads due to

a near field detonation. Simulation of highly non-linear material behaviour is a challenging task. Strain rate dependency of material properties, as described in Section 3, needs to be taken into consideration. For extremely large strain rates (shock propagation), there are several non-linear material models available.

Hydrodynamic material models using Mie–Grüneisen equations of state can be used. The SPH approach provides one possibility of simulating the breach through a structure. One numerical simulation study with experimental results is presented and discussed in Ref. [73].

Analyses with these sophisticated techniques are usually performed when simpler methods, generally with unquantifiable conservatisms, are likely to lead to conclusions of adverse failures. Other likely applications are for benchmarking simpler methods, especially when dealing with complex geometry or not easily modelled situations, such as explosions in confined or partially confined spaces.

Finite element methods for explosion loads are not further developed in this report; the focus is on simplified methods. Strictly speaking, simplified methods are not just calculating the response. Owing to the nature of the problem, elements of the capacity evaluation need to be introduced to permit their application.

FIG. 29. Real and equivalent structural systems.

4.3.3.2. Development of an equivalent single degree of freedom system

The mass and dynamic load of the equivalent SDOF system (Fig. 30) are based on the component mass and blast load, respectively; the spring stiffness and yield load are based on the component flexural stiffness and load capacity.

The properties of the equivalent SDOF system are also based on load and mass transformation factors. These factors are calculated such that the SDOF system and the represented component have equal kinetic, work and strain energies at each time, assuming that the SDOF system deflection is always equal to the component maximum deflection and that the component maintains an assumed deformed shape as it responds to a blast load.

The basic equation of motion for an SDOF system under blast load is as follows:

e e( ) e( )

M x R x+ =F t (59)

where

Me is the effective mass;

Re is the stiffness dependent effective resistance function;

and Fe(t) is the effective blast load history.

M

x F (t)

K

F (t) F

td t

FIG. 30. A single degree of freedom elastic structure subject to an idealized blast pulse.

Comparing with the basic equation motion for an SDOF system under general dynamic loading, it should be noted that for blast loading the damping term can be ignored.

Under blast loading, the equivalent SDOF system deflects beyond the yield deflections and the formation of plastic hinges changes the deflected shape. This is taken into account using transformation factors and Eq. (59) becomes:

M c R c( ) L c( )

K M x K R x+ =K F t (60)

where

KM is the mass transformation factor;

KR is the resistance transformation factor;

KL is the load transformation factor;

Mc is the mass of the blast loaded element;

and Rc is the resistance of the blast loaded element.

It can be shown that the resistance factor KR has to always equal the load factor KL [7]. This equation can then be simplified if the load and mass transformation factors are combined into a single load mass factor KLM:

LM c c( ) c( )

K M x R x+ =F t (61)

where KLM = KM/KL.

This equation is solved to determine the deflection history of the equivalent SDOF system, which is equal to the maximum defection history of the blast loaded element.

Basically, there are three different regimes regarding the structural behaviour under blast loading, which correspond to three different types of solution for Eq. (61):

(a) The positive phase duration is long compared to the natural period of vibration of the structure. In this case, the load may be considered as being constant while the structure attains its maximum deflection. This can be a case derived from a very strong blast source at a great distance. Such loading is referred to as quasi-static or ‘pressure’ loading.

(b) The positive phase duration is short compared to the natural period of the structure. In this case, the load has finished acting before the structure has had time to respond significantly. The maximum deformation of the

structure occurs well after time td in Fig. 30. In this case, the structure is subjected to ‘impulsive’ loading (i.e. a momentum is transferred to the system, which translates into an initial velocity of the mass M in Fig. 30).

(c) The positive phase duration is similar to the natural period of the structure.

In this case, the assessment of the response is more complex, possibly requiring a complete solution of the equation of motion (as mentioned in Eq. (59), the damping term is neglected), though it is often possible to have a reasonable approximation to the response by using results obtained for impulsive or quasi-static loading. In this case, the structure is said to be subjected to ‘dynamic’ loading.

The entire design process using SDOF systems is shown in the flow chart in Fig. 31. Examples of calculations illustrating the design of structures for blast loads are presented in Appendix III.

Define the simplified SDOF system as a function of the structure or structural element which has to be  designed and determine its mass per unit area loaded by blast. 

Determine blast load on the structure or structural element. 

Determine the flexural stiffness of structured or structural element considering yielding in maximum  moment region as well as its resistance­deflection relation based on resistance and stiffness values

Define equivalent SDOF equation as a function of the response regime of the structure (impulsive,  dynamic, and quasi­static) and solve the equation for the maximum dynamic value of studied parameter 

(e.g. deflection). 

Convert calculated maximum deflection to support rotation/ductility and compare to allowable criteria. 

Determine maximum equivalent static reactions and design the structure or element for shear. 

FIG. 31. Design process using single degree of freedom systems.

4.3.3.3. Pressure–impulse diagrams

Pressure–impulse diagrams are used in blast resistant design for quick assessment of structural elements or structural systems. This approach takes advantage of the non-dimensional scaling characteristics of the phenomenon to define the onset of structural damage as a function of the pulse loading parameters (Fig. 32).

The information necessary to construct a pressure–impulse diagram includes the equivalent static and dynamic characteristics of the structure being evaluated (i.e. those needed to build an SDOF equivalent system) and the shape of the loading pulse (e.g. triangular, rectangular). Then, for each pair of peak pressure Pmax and impulse of the pressure distribution I, the maximum deflection is computed using an SDOF system. The computed maximum deflection is used to assess the expected level of damage.

It should be noted that once the peak pressure and pulse shape are fixed, pulse duration is proportional to the given impulse. Hence, for large impulses, the response will enter into the quasi-static regime, where the maximum dynamic deflection is equal to a constant factor times the static deflection (the maximum amplitude of load P divided by the equivalent stiffness K). For elastic systems, this factor is equal to two. For inelastic systems, the factor is less than two. As discussed earlier, this regime corresponds to loading functions where the time

101 102 103

100 10-2 10-1

100 101

Impulse is (kPa. S)

Pressure PS (kPa)

(A) Severe damage

(B) No damage / minor damage

FIG. 32. A typical pressure–impulse diagram.

of application is long compared to the fundamental period of the structure.

Once within this regime, the maximum deflection is dependent only on the stiffness characteristics of the system and the amplitude of the loading function (i.e. independent of the mass of the system and the total duration of the pulse).

On the other hand, once the impulse and the pulse shape are fixed, pulse duration is inversely proportional to peak pressure. Hence, for large peak pressures, the response will enter into the impulsive regime. As discussed earlier, this regime corresponds to loading functions where the time of application is short compared to the fundamental period of the structure. Once within this regime, the maximum deflection is dependent only on the impulse value and the dynamic characteristics of the system.

Consequently, the part of the pressure–impulse diagram corresponding to these two first regimes of the response can be constructed relatively easily.

A third regime is intermediate to the other two. It is the dynamic loading regime. In this regime, numerical analysis needs to be employed to calculate the maximum response. Again, however, it is relatively straightforward to do so, once the loading function is known and the equivalent SDOF representation of the structure or structural element has been built.

Failure is most often defined by deformation limits, which allows the evaluation to account for the significant effect of non-linear energy absorption and ductility. For analysis using pressure–impulse diagrams, curves of equal deformation, in terms of ductility factors can be developed in order to facilitate the work of the engineer making the assessment.

4.4. STRUCTURAL RESPONSE ANALYSIS FOR THERMAL LOADS

문서에서 S a f e t y R e p o r t s S e r i e s N o. 8 7 (페이지 101-107)