Добавил:
Upload Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Учебники / 0841558_16EA1_federico_milano_power_system_modelling_and_scripting

.pdf
Скачиваний:
72
Добавлен:
08.06.2015
Размер:
6.72 Mб
Скачать

188 8 Time Domain Analysis

¯

Under certain hypothesis, it is possible that the admittance matrix Y does not depend on state variables. For example by modelling dynamic series devices (e.g., regulating transformers or FACTS devices) as current injections at the sending and receiving buses, respectively. In this case the only links among dynamic devices are the algebraic variables v¯ through the admittance

¯

 

 

matrix Y . For this reason, v¯ are also sometimes called aggregation variables.

Constant admittance loads allows simplifying (8.13)

since, for a constant

¯

2

Thus,

admittance, the correspondent element of the vector i(x, v¯) is zero.

ordering the vector of bus voltages into a vector of generator bus voltages v¯G and a vector load bus voltages v¯L, the algebraic equations in (8.13) becomes:

¯

¯

 

¯

0 = iG(x, v¯) + Y GGv¯G + Y GLv¯L

¯

¯

 

 

0 = Y LGv¯G

+ Y LLv¯L

 

where:

¯

¯

 

Y =

Y LG Y LL

¯

Y GG Y GL

 

¯

¯

 

Thus, load bus voltages can be eliminated from (8.14):

 

 

¯ 1 ¯

 

 

 

v¯L = −Y LLY LGv¯G

 

¯ 1 ¯

 

¯

¯

¯

0 = iG(x, v¯) + (Y GG − Y GLY LLY LG)v¯G

¯

Finally, by defining a reduced admittance matrix Y r as:

¯ ¯ ¯ ¯ 1 ¯

Y r = Y GG Y GLY LLY LG

the system (8.13) becomes:

x˙ = f (x, v¯G)

 

¯

¯

0 = iG(x, v¯) + Y r v¯G

(8.14)

(8.15)

(8.16)

(8.17)

(8.18)

The latter model is the most commonly used in transient stability analysis, especially in proprietary software packages. The advantage of this formulation is that the order of algebraic equations is consistently reduced with respect to the full system size since generator buses are much less than load and transit nodes. Furthermore, algebraic equations are linear and most elements of the reduced admittance matrix are constant (although they can vary due to line outages, fault occurrences and load shedding).

The current-injection model (8.18) is a standard de facto for transient stability analysis. However, any model, even the most well-accepted one, is

2Observe that pure transit nodes are a special case of loads with a zero constant admittance. In the following, pure transit nodes are implicitly considered constant admittance loads.

8.2 Power System Model

189

simply a model, subjected to hypothesis. The most restrictive hypothesis that leads to (8.18) is to assume constant impedance loads. This hypothesis is reasonable only if the time frame is that of transient stability (e.g., few seconds following a short circuit occurrence). In fact, load controls (e.g., tap changer voltage regulation) can be considered frozen for the few seconds following a large disturbance. On the other hand, voltage and frequency stability analyses require detailed models of dynamic loads and their controls [146]. Constant impedance loads are also inadequate for long-term voltage and frequency stability analyses. As a matter of fact, for long-term analysis, tap changer voltage control allows modelling loads as constant power consumptions.

8.2.2Power-Injection Model

The power injection mode is obtained from (8.13) by multiplying the conjugate of algebraic equations by bus voltages:

 

 

x˙ = f (x, v¯)

 

 

 

 

 

 

 

 

(8.19)

 

 

¯ ¯

(x, v¯)

 

 

¯ ¯

 

(x)v¯

 

 

 

 

0 = V i

V Y

 

 

 

 

 

 

 

 

 

 

 

 

¯

 

 

 

 

 

 

 

¯ ¯

(x)v¯

are the power flow

where V = diag(¯v1

, v¯2, . . . , v¯nb ). The term V Y

¯ ¯

(x, v¯) are the complex powers injected at network buses.

equations, while V i

 

Thus, (8.19) can be rewritten as:

 

 

 

 

 

 

 

 

 

 

 

x˙ = f (x, v¯)

 

 

 

 

 

 

 

 

(8.20)

 

 

0 = s¯(x, v¯)

¯ ¯

(x)v¯

 

 

 

 

V Y

 

 

 

 

 

 

 

 

 

 

 

 

 

Equations (8.20) are equivalent to (8.13) but are intrinsically nonlinear and are thus computationally more demanding than (8.13).

Another issue of writing algebraic equations in terms of power injections is numerical. Assume that, as a consequence of a short-circuit, some bus voltage magnitudes become zero. After clearing the fault, voltage magnitudes recover positive values. However, if one uses a Newton’s method to solve algebraic equations, one has, at a certain bus h where vh = 0:

0 = vheh ¯ih(x, v¯)

n

y¯hkv¯k = vhκ(z)

(8.21)

 

 

 

 

k=1

where z = [xT , vT , θT ]T . The Newton’s equation for (8.21) at a generic iteration i is:

vh

κ(z(i)) = κ(i)(z(i))Δvh

+ vh

 

nz

∂z

(i)

(i)

 

 

∂κ(i)

 

 

 

 

 

 

 

 

 

 

=1

 

Δz(i) (8.22)

190 8 Time Domain Analysis

If v(i)

= 0 at a certain iteration i, then (8.22) becomes:

 

h

 

 

 

0 = κ(i)(z(i))Δv(i)

(8.23)

 

h

 

which lead to Δvh(i) = 0. In other words, if vh(i) = 0 at the iteration i, it

will remain zero for all the following iterations. A very small value of vh(i) would also show a similar numerical issue. Furthermore, v = 0 is a solution of algebraic equations in (8.20). Thus, one has to carefully avoid that voltages become zero (or very small values) at any iteration, otherwise, the Newton’s method is not able to recover voltage magnitudes.

In (8.20), the only algebraic variables are bus voltage magnitudes and phase angles. A more general and flexible model includes additional algebraic variables yˆ, algebraic equations gˆ and controllable parameters η:

x˙ = f (x, yˆ, v¯, η)

 

 

(8.24)

0 = gˆ(x, yˆ, v¯, η)

 

 

 

0 = s¯(x, yˆ, v¯, η)

¯ ¯

(x, yˆ, η)v¯

V Y

 

 

 

The following conclusive remarks are relevant:

1.Both (8.13) and (8.24) are nonlinear. In fact, (8.13) is nonlinear at least in the di erential equations of synchronous machines. Thus, if an implicit solution method is used for the numerical integration, both (8.13) and (8.24) requires a Newton’s method for solving a set of nonlinear equations at each integration step.

2.Equations (8.24) have the advantage of requiring the same model for power flow, continuation power flow and time domain analyses. In other words, (8.24) allow using an unique structure for all devices (there is no di erence between power flow and time domain analysis models of the same device) and writing more compact code (the same algebraic equations are used for both static and dynamic analysis).

Example 8.1 OMIB Di erential Algebraic Equations

This example provides the power-injection as well as the current-injection model (8.24) and (8.13), respectively, for the 2-bus system depicted in Figure 8.6. Bus 0 is an infinite bus where the voltage v¯0 = v0 θ0 is constant, while the machine is a two-order machine model. Assuming that the voltage at the infinite bus is a parameter, the system power-injection model is:

f

ω˙ = (pm

 

pe D(ω

 

ωs))/2H

(8.25)

 

˙

 

 

ωs)

 

 

 

 

δ = ωn(ω

 

 

 

 

 

 

 

8.2 Power System Model

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

191

 

 

 

 

1

 

 

 

 

 

 

 

 

 

 

 

 

0

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

xL

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

v1 θ1

 

 

 

 

 

 

 

 

 

v0 θ0

 

 

 

 

 

 

 

 

 

Fig. 8.6

 

OMIB system

 

 

 

 

 

 

 

0 = vqq+ raiq

 

 

 

eq

+ xd

id

 

 

pe

 

 

 

0 = (v

+ raiq )iq

+ (vd

+ raid)id

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

gˆ

 

0 = v + r i

d

 

 

 

x i

q

 

 

 

 

 

 

 

 

 

 

 

d

a

 

 

 

 

d

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

0 = v1 sin(δ

 

 

θ1)

 

 

vd

 

 

 

 

 

 

 

 

 

0 = v1 cos(δ

θ1)

vq

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

v1v0

sin(θ1

 

θ0)

 

 

 

 

 

0 = vdid + vq iq

 

 

 

xL

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

θ0)

s¯ 0 = vq id vdiq

v1 + v1v0 cos(θ1

 

 

 

 

 

 

 

 

 

2

 

 

 

 

 

 

 

 

 

 

 

 

 

xL

 

 

xL

 

 

 

 

The current-injection model can be obtained by substituting s¯ in (8.25) with the current injection:

¯

 

1

 

 

 

iG

 

0 = (id + jiq )

jxL

v1

− v¯0)

(8.26)

In (8.26), the generator current id + jiq can be substituted by explicit functions of x and v¯1. From (8.25), one obtains:

i

 

 

1

x

r

 

e

v

 

iq

xd2 + ra2

ra

−xd

−vd

 

d

=

 

d

a

 

q

q

(8.27)

 

 

 

 

 

 

where vd and vq are functions of the rotor angle δ and of the bus voltage v¯1. Assuming ra 0, which is a common hypothesis for the classical machine model, (8.27) can be further simplified as:

i

d

= (e

v )/x

= (e

v cos(δ

θ

))/x

 

q

q

d

 

q

1

1

d

i

q

= v /x

=

v

sin(δ

θ

)/x

 

 

 

 

d d

 

1

 

1

d

 

 

 

and the electrical power pe becomes:

eq v1

pe = xd sin(δ − θ1)

In summary, the simplified current-injection model is:

f

˙

 

eq v1

sin(δ

θ1)

D(ω

 

ωs))/2H

ω˙ = (pm

 

 

 

δ = ωn(ω

 

ωs)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

¯iG

 

 

xd

 

 

 

-0 = xd (eq

− v1 cos(δ − θ1) + jv1 sin(δ − θ1)) jxL v1

 

1

 

 

 

 

 

 

 

 

 

1

 

(8.28)

(8.29)

(8.30)

− v¯0)

192

8 Time Domain Analysis

8.3Numerical Integration Methods

In order to numerically integrate (8.2), the first issue that has to be solved is how to handle algebraic equations g. There are mainly two approaches:

1.Partitioned-solution approach. Variables x and y are updated sequentially.

2.Simultaneous-solution approach. Variables x and y are solved together in a unique step using a solver such as the Newton’s method.

As usual, both approaches have advantages and drawbacks.

In the partitioned approach, since x and y are updated independently, any numerical integration method can be used. However, the sequential approach is typically used combined with explicit numerical methods (e.g., RungeKutta’s formulæ) that do not require computing and factorizing the Jacobian matrix f x. On the other hand, the partitioned approach introduces a “delay” between x and y. In fact, for a generic step i, while computing x(i+1), algebraic variables are frozen to the old value y(i). Moreover, the state variables x(i+1) are not modified when computing y(i+1). To avoid the delay between x(i) and y(i), one has to iterate over x(i) and y(i) for each time step. This process can lead to numerical instabilities. It has to be noted that solving g = 0 for updating algebraic variables requires the solution of a nonlinear system, which generally requires computing and factorizing iteratively the Jacobian matrix gy . At this aim, to reduce the computational e ort, one can use a Newton’s dishonest method as described in Subsection 4.4.6 of Chapter 4. In case the Jacobian matrix of algebraic equations is kept constant for multiple integration time steps, the method is called very dishonest Newton’s method [19].

The simultaneous approach has the advantage that x(i) and y(i) are updated together, thus no delay is introduced. This approach is used in conjunction with implicit numerical methods that require, at each time step, the solution of a set of nonlinear equations. This solution is generally obtained through a Newton’s method. Thus, the simultaneous approach requires iteratively computing and factorizing an (nx + ny) × (nx + ny) Jacobian matrix.

In conclusion, the partitioned approach can be considered faster but less numerically stable than the simultaneous approach. Further details are given in the following subsections.

8.3.1Explicit Methods

Multi-stage explicit methods can be expressed using a m-stage formula in the form [276]:

8.3 Numerical Integration Methods

 

193

x(t + Δt) = x(t) + Δt

m−1

(8.31)

ckf {k}

 

 

 

 

k=0

 

xˆ(t + Δt) = x(t) + Δt

m

 

cˆk f {k}

 

 

 

 

 

k=0

 

where t is the current integration time and Δt is the step length and:

 

f {0} = f (x(t), ti)

 

(8.32)

k−1

 

 

f {k} = f (x(t) + Δt bkj f {j}, ti + akΔt)

 

 

 

 

j=0

Both x(t + Δt) and xˆ(t + Δt) approximate the exact solution and xˆ(t + Δt) is an approximation of higher order than x(t + Δt). The di erence x(t + Δt) − xˆ(t + Δt) allows estimating the error with respect to the exact solution and adjusting the step length Δt.

The m-stage formula (8.31) evaluates the new value of the state variables x(t + Δt) using a weighted sum of m values of x˙ at suitable points between t and t + Δt.

A convenient way of visualizing the formulæ (8.31) for a given method is through the Butcher’s tableau [34], as follows:

0

 

 

 

 

 

 

 

a1

b10

 

 

 

 

 

 

a2

b20

b21

 

 

 

 

 

.

. . .

.

 

(8.33)

.

.

.

 

.

.

. .

 

 

 

am

bm0 bm1 . . . bm,m−1

 

 

 

c0

c1

. . . cm−1

 

 

 

cˆ0

cˆ1

. . . cˆm−1

cˆm

Using (8.31), a huge variety of explicit methods can be defined, including the large family of Runge-Kutta’s formulæ.

Multi-step or predictor-corrector methods are another class of explicit methods. The general formula of a multi-step method is:

m

 

 

 

x(t + Δt) = x(t) + Δt ck f (x(t − kΔt), t − kΔt)

(8.34)

k=0

A drawback of multi-step methods is that they are not self-starting, since the first m steps has to be known to compute the generic step (8.34) for t+Δt. The well-known Adams-Bashforth’s method, Milne-Simpson’s method and Hamming’s method belong to the family of multi-step methods [34]. However, these methods have proved to be less accurate and e cient then Runge-Kutta’s formulæ, at least for power system applications [163].

194

8 Time Domain Analysis

Example 8.2 Runge-Kutta’s Formulæ

Example 4.6 presented the classical 4th order Runge-Kutta’s formula:

f {0} = f (x(t))

(8.35)

f {1} = f (x(t) + 0.5Δtf {0}) f {2} = f (x(t) + 0.5Δtf {1}) f {3} = f (x(t) + Δtf {2})

x(t + Δt) = x(t) + Δt(f {0} + 2f {1} + 2f {2} + f {3})/6

The RK4 is represented by the following Butcher’s tableau:

0

11

22

1

0

1

 

(8.36)

2

 

2

 

 

 

1

0 0 1

 

 

 

 

 

 

 

 

 

1

2

2

1

 

 

6

6

6

6

 

Several more sophisticated schemes have been proposed [117]. For example, the Runge-Kutta-Fehlberg’s formula has the following Butcher’s tableau:

0

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1

1

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

4

4

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

3

 

3

 

 

 

 

9

 

 

 

 

 

 

 

 

 

 

8

32

 

 

 

 

32

 

 

 

 

 

 

 

 

 

 

12

1932

7200

7296

 

 

 

 

 

 

 

 

13

2197

2197

2197

 

 

 

 

 

(8.37)

1

216439

 

 

8

3680513

845

 

 

 

 

 

 

4104

 

 

 

 

 

 

21

8

 

 

 

2

 

25653544

41041859

4011

 

 

 

27

 

 

 

 

 

 

25

 

 

 

 

0

 

1408

2197

1

 

 

 

 

 

 

 

 

 

 

 

 

 

5

 

 

 

 

216

 

 

 

 

2565

4104

 

 

 

 

16

 

 

 

 

0

 

6656

28561

9

 

2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

135

 

 

 

 

12825

56430

50

55

 

 

Example 8.3 Modified Euler’s Method

The modified Euler’s method is the simplest multi-step method and is the only that has been widely used in power system analysis [163, 291]. It is composed of a predictor and a corrector step, as follows:

8.3 Numerical Integration Methods

195

x˜(t + Δt) = x(t) + Δtf (x(t), t)

(8.38)

x(t + Δt) = x(t) + 12 Δt(f (x(t), t) + f (x˜(t + Δt), t))

The accuracy of this method can be improved by assigning x(t) ← x(t + Δt) and repeating the two-step formula (8.38).

8.3.2 Implicit Methods

In transient stability analysis, one of the possible issues is that time constant can span various time scales. For example if both transient stability and long term dynamics are considered together, time constants vary between 102 and 103 s (see Figure 1.6 of Chapter 1). Similarly, if one considers both sub-synchronous resonance phenomena and transient stability, the time scale range is between 104 and 101 s. If the time scale range of a ODE problem is “big”, the ODE problem is said to be sti .

The behavior of numerical methods on sti ODE problems can be analyzed

by applying these methods to the test equation:

 

x˙ = kx, k C

(8.39)

The solution of (8.39) is x(t) = ekt that approaches zero as t → ∞ when{k} < 0, i.e., the left-half of the complex plane is the stability region of the test equation (8.39). If the numerical method also exhibits this behavior, then the method is said to be absolute stable or A-stable according to the Dahlquist’s definition [69]. For example, applying the Runge-Kutta’s formulæ to the test equation (8.39), one has:

x(t + Δt) = χ(kΔt)x(t) = χn(kΔt)x(t0)

(8.40)

where n is the number of steps and x(t0) is the initial value. The function χ(kΔt) is called stability function and must be (kΔt)| < 1 to satisfy x(t) 0 for n → ∞.

An interesting result of Dahlquist’s theorems is that an explicit multistep method cannot be A-stable [69, 165, 316]. Thus, explicit methods are expected to provide poor behavior for sti ODE problems. On the other hand, implicit methods can be A-stable. For this reason, and because implicit methods allow a simultaneous solution of both state and algebraic variables in DAE problems, implicit methods are of particular relevance for power system analysis.

When using implicit methods, each step of the numerical integration is obtained as the solution of a set of nonlinear equations. Thus, implicit methods are particularly suited for nonlinear DAE systems, since the algebraic equations can be included in the nonlinear system to be solved at each iteration.

For a generic time t, and assumed a step length Δt, one has to solve the following problem [30]:

196

8

Time Domain Analysis

 

0 = qˆ(x(t + Δt), y(t + Δt), f (t))

(8.41)

 

0 = g(x(t + Δt), y(t + Δt))

 

where f and g are the di erential and algebraic equations and qˆ is a function that depends on the implicit numerical method. Equations (8.41) are nonlinear and their solution is obtained by means of a Newton’s method, which in turn consists of computing iteratively the increments x(i) and y(i) of the state and algebraic variables and updating the actual variables:

 

x(i)

=

(i)

 

qˆ(i)

 

 

 

 

 

y(i)

[Ac

]1

g(i)

 

 

(8.42)

x(i+1)

(t + Δt)

 

x(i)(t + Δt)

 

 

x(i)

 

y(i+1)

(t + Δt) =

y(i)(t + Δt) +

y(i)

where A(ci) is a matrix depending on the algebraic and state Jacobian matrices of the system. Some examples of implicit method formulæ are given in the following examples. Most techniques described in Chapter 4 can be applied to (8.42). For example, the dishonest or very dishonest Newton’s methods can be useful to reduce the number of factorizations of A(ci) and speed up the simulation.

Example 8.4 Backward Euler’s Method

The backward Euler’s method is a first order implicit method. It is generally faster but less accurate than the trapezoidal method that is discussed in the next example. At a generic time step t + Δt and a generic iteration i, A(ci)

ˆ(i)

and q are as follows:

c

!

gx(i)

gy(i)

"

A(i)

=

Inx − Δtf x(i) −Δtf y(i)

(8.43)

qˆ(i)

= x(i)(t + Δt) − x(t) − Δtf (i)

where Inx is the identity matrix of the same dimension of the dynamic order

of the DAE system and all Jacobian matrices and f (i), are computed at the current point (x(i)(t + Δt), y(i)(t + Δt), t + Δt).

Example 8.5 Trapezoidal Method

The Crank-Nicolson’s or trapezoidal method is the workhorse solver for electro-mechanical DAE, and is widely used, in a variety of flavors, in most commercial and non-commercial power system software packages. The implicit version of the trapezoidal method has proved to be very robust and

8.3 Numerical Integration Methods

197

reliable for a variety of sti ODE and DAE systems. At a generic iteration i,

A(i) and qˆ(i)

are as follows:3

 

 

c

 

!

gx(i)

gy(i)

"

 

c

 

A(i)

=

Inx 0.5Δtf x(i) 0.5Δtf y(i)

(8.44)

 

qˆ(i)

= x(i) − x(t) 0.5Δt(f (i) + f (t))

where the notation is the same as in (8.43).

Example 8.6 Rosenbrock’s Semi-Implicit Method

Dahlquist’s theorems discourage from setting up implicit methods of order greater than 2 [165]. However, an interesting method for improving the accuracy of an implicit method is proposed in [316]. This reference proposes semi-implicit methods for solving ODE systems which are A-stable and avoid the need of iterating. However, semi-implicit methods applied to nonlinear DAE systems still require iterating and factorizing the matrix A(ci) at each iteration.

A well-known semi-implicit method is the one based on Rosenbrock’s formulæ. For example, a 4th order Rosenbrock’s formula is as follows. At a generic time t and iteration i, the matrix A(ci) is the same as in (8.44). The variables x(i) and y(i) are determined by means of a 4th order approximation:

qˆ(i)

= [

1

 

A(i)]1

ϕ(t)

 

 

(8.45)

0.5Δt

 

 

1

 

c

 

 

 

 

 

 

 

 

 

 

 

z1(i) = z(t) + a21qˆ1(i)

 

 

 

 

qˆ(i)

= [

1

 

A(i)]1

(ϕ(t + a2xΔt) + c21qˆ(i)

/Δt)

 

 

 

 

0.5Δt

 

2

 

c

1

 

 

 

 

 

 

 

 

 

 

z2(i) = z(t) + a31qˆ1(i) + a32qˆ2(i)

 

 

 

qˆ(i)

= [

1

 

A(i)]1

(ϕ(t + a3xΔt) + (c31q(i)

+ c32qˆ(i))/Δt)

0.5Δt

3

 

c

1

 

2

 

qˆ(i)

= [

1

 

A(i)]1

(ϕ(t + a3xΔt) + (c41qˆ(i)

+ c42qˆ(i)

+ c43qˆ(i))/Δt)

0.5Δt

4

 

c

1

 

2

3

 

 

 

 

 

 

 

z(i) = z(t) + b1qˆ(1i) + b2q(2i) + b3qˆ(3i) + b4qˆ(4i)

3 Butcher’s tableaux can be defined also for implicit methods. The only di erence with explicit methods is that the coe cient matrix bkj is not necessarily lower triangular. For example, the Butcher’s tableau of the trapezoidal method is:

1 1 1

2 2

0 0 0

1 1

2 2

However, for implicit methods, the determination of each f {k} is generally involved and, hence, the Butcher’s tableau representation is not practical.