Conference PaperPDF Available

Simulating Design Processes with self-iteration activities based on DSM planning

Authors:

Abstract and Figures

Design processes are fundamentally iterative. The increasing knowledge about the product, while designing, manifests in design changes of previously accomplished activities. Such iterations are considered as a major source of increased product development lead-time and cost; thus process planning and simulation are essential. The design structure matrix (DSM) is a known method for process planning. Since the DSM itself does not express all the relevant information that is required for defining process logic, process logic interpretation is required. "Business rules", being logic interpretation options that are applicable in different business cases, can guide automatic translation of the DSM to valid iterative process plans. This study focuses on implementing self activity iterations in process simulation. Notwithstanding their importance, they are overlooked in DSM literature. Using different business rules results in various run time processes that exhibit non-trivial behaviors, interpretation rules are fully validated for a test case.
Content may be subject to copyright.
Simulating Design Processes with self-iteration activities based on DSM planning
Arie Karniel
ariek@eng.tau.ac.il
Yoram Reich
yoram@eng.tau.ac.il
School of Mechanical Engineering, Tel Aviv University, Tel Aviv, Israel
Abstract
Design Processes are fundamentally iterative. The
increasing knowledge about the product, while
designing, manifests in design changes of previously
accomplished activities. Such iterations are considered
as a major source of increased product development
lead-time and cost; thus process planning and
simulation are essential.
The Design Structure Matrix (DSM) is a known
method for process planning. Since the DSM itself does
not express all the relevant information that is required
for defining process logic, process logic interpretation
is required. “Business Rules”, being logic
interpretation options that are applicable in different
business cases, can guide automatic translation of the
DSM to valid iterative process plans.
This study focuses on implementing self activity
iterations in process simulation. Notwithstanding their
importance, they are overlooked in DSM literature.
Using different business rules results in various Run
time processes that exhibit non-trivial behaviors.
interpretation rules are fully validated for a test case.
1. Introduction
Product Development Processes (PDP) can be
defined as collection of related activities targeted to
converting a new concept (market opportunity) into a
marketed product. Within PDP, the cyclic Design
Process (DP) incorporates synthesis, analysis, and
decision-making stages to migrate from problem
definition to artifact specification that meet the
marketing and business case vision. The Design
Process reflects the specific design requirements,
objectives and constraints applicable to the product.
Iterations are inherent to DPs, as changes in the
design of one component may result in changing the
design of other components. Additionally, the process
includes testing activities that are a-priori planned
iterations [14] and unpredicted feedback rework due to
unplanned changes or problems. Iterations are
considered as major source of increased product
development lead-time and cost [6] [9].
One type of iteration occurs when an activity is
iterated several times before completion. Self-iteration
of a design activity could always be separated to
design, check and decide. Such distinction, (and
subsequently, proliferation in the Design Structure
Matrix (DSM)), might be justified if the distinct
activities are performed by different resources, but
might be just an overhead if done by the same designer
or design resource. If all iterations are completed
before continuing to the next activity, the only impact
of self iteration might be duration extension. However,
if such self-iteration is executed in parallel to other
activities, its influence span may extend to all linked
activities that are executing or completed execution
between the completion of last iteration and the current
one. Self activity iterations are briefly mentioned in
[10], but the implications of including activity self
iterations within concurrent process implementation,
are somewhat overlooked in the DSM-based simulation
literature. The inclusion of self iterations, elicits a wide
range of potential Run Time process structures. Some
of the additional process characteristics are manifested
in the cases of overlapping parallel activities [7].
Planning the process and performing project
activities in an appropriate sequence is critical in order
to minimize rework due to information dependencies.
Introduced by Steward [8], the DSM is utilized to
capture the required product knowledge for process
planning. Yet, the DSM does not represent any
information regarding the linkages logic, and the
presented information is insufficient for implementing
process simulations [8] [23].
Process verification issues being thoroughly
discussed in process modeling literature [2] [18], are
seldom mentioned in the DSM literature [12].
Verification checks are essential for automatic
translation of DSM to concurrent process plan model,
the Process Scheme. Translating the DSM is not
unique due to interpretation options based on the
applicable business cases [13]; these interpretations
defined as Business Rules may impact the process
scheme or merely the actual Run Time processes.
In what follows, section 2 surveys the DSM and
process verification literature and section 3 merges
them into the verification of DSM-based process plans.
Section 4 analyzes implementing the method from
section 3 to the case of self iterations for a simple case
study. This analysis elucidates the behavioral richness
and the complexity of implementing self-iterations.
Conclusions are drawn in section 5.
2. Literature survey
The term Activity is used to describe a logic step
within a process. Activities have pre and post
conditions; additionally, they may accept inputs and
share outcomes during their performance [7]. The term
Process Scheme, is used to describe the process
structure, i.e., the model of linkages between activities,
their precedence, concurrency, and logic relations [12].
The term Run Time process, is used to describe the
process progress according to the Process Scheme.
A limited set of process constructs are being used to
define the Input logic of an activity (pre-conditions)
and its Output logic (post conditions). Karniel and
Reich [12] used the construct definitions (Workflow
patterns) elaborated in [3]. The constructs utilized in
this study are
1
: Serial link {1}; Split-And {2}, sending
a termination signal through all output links; Split-Or
{6}, sending termination signal to some of the output
links; Split-Xor {4}, sending to only one of the output
links; Join-And {3}, wait for all input signals; Join-Or
{8}, wait for first to come, execute multiple times; and
Join-Xor {5}.
2.2. Process planning using DSM
The commonly used GANTT and Pert charts are
inadequate for planning design processes, as they do
not effectively model design activities dependencies
and process iterations [23]. The DSM representation
can be used for various applications [5]. The DSM
provides the means to identify serial, parallel and
iterative design flows; model and manipulate iterations
and multidirectional information flows [9]. DSM is
square matrix that uses off-diagonal entries to signify
the dependency of one element on another. When
modeling processes, the lower diagonal portion
represents a precedent activity relationship; i.e., a
marking in cell {j,k} (row j, column k, j>k), indicates
that activity k should follow activity j. For each
activity, its row shows its inputs and its column shows
its outputs. The upper diagonal portion of the DSM
matrix denotes an iterative process (a link to an
upstream activity). The outcome of a particular activity
can directly impact a previously performed activity,
which should be performed again, i.e., rework.
Serial activities are linked by forward link, and
indicate a serial process, Figure 1(a). Parallel or
concurrent activities have no relation linkages, Figure
1(b). Coupled activities have forward and feedback
links, Figure 1(c). Additional relation type, Overlap
1
The numbers in bracket indicate the definitions in [3].
Synchronizing Merge {7} is not used in the current study
.
activities can be considered as a Parallel relation [24].
Cho and Eppinger [7] model two overlap activities as
parallel activities with additional lead time, and a
constraint preventing completion of the latter activity
before the first activity. Coupled activities form an
activity loop (cycle). When the cycle includes three
activities (and more), the cycle may have many forms
in the DSM representation [12].
Different DSM markings are typically used.
Binary DSM uses tick-marks (‘x’) to represent
precedence relations [22], (e.g., Figure 1). Numeric
DSM represents a measure of the degree of relation or
its importance ranking. Probability DSM [21] is used
in this study. Probabilities are assigned directly or
result from transformation of Numeric (or Binary)
DSM [19] [23]. The reordering algorithms: Partitioning,
Sequencing, and Clustering typically do not utilize
diagonal values. The diagonal values are typically left
empty, or used for describing activity properties (e.g.
Duration). In this study the diagonal values are used
for self-iteration probabilities.
Figure 1. Activity relations
The information presented in the DSM is
insufficient for simulation, e.g., activity duration. Such
information should be specified in addition to the
DSM. Since such information is not included,
algorithms that use only the DSM structure, are
criticized as inadequate for optimization, and require
simulation-based optimization [6].
Karniel and Reich [12] surveyed various simulation
process structures and process progress types,
including Deterministic progress, Markov chain or
Monte Carlo simulations. The survey analyzed the
translation of DSM-based process plan, into process
simulation in various studies, classified the approaches
used, and discusses their strengths and limitations
along problems related to process modeling
verification.
2.3. Process verification
Process correctness verification is profoundly
discussed in the Workflow literature [16] [18]. An
established formal languages model for process
specification is Petri net [15], which is a bipartite
directed graph with two node types called places and
transitions, connected by directed Edges (arcs). Places
can contain Tokens. A transition may transfer tokens
from its input places to its output places according to
rules. Process activities are modeled by transitions;
places correspond to conditions. Marking, representing
the process state, is the distribution of tokens over the
places. Petri nets can be used as diagrammatic tool for
modeling and analyzing distributed system behavior
including sequential, conditional, parallel and iterative
routing [2]. WF nets (workflow nets) [1] are Petri nets,
used for specifying the control of workflow processes
with defined starting and ending. Using the formal
definitions of a WF net one can set the required
conditions for proper process specification (Process
Scheme), using a correctness criteria named
‘Soundness’. WF net properties are [1]: 1) A WF net
has starting place and final place. 2) All the activities
and places are on a path from the starting place to the
final place (i.e., no activities or conditions that do not
contribute to process progress).
Proper process, defined according to the soundness
criteria has the following properties. 1) From every
process state, which is reachable from the initial state,
(token at the starting place); there is a ‘firing sequence’
that leads to termination state. The process should
terminate eventually. 2) Once the terminal state was
reached there are no activities that did not perform yet;
formally: there are no tokens in places other than the
final place (i.e. the termination state). 3) There are no
inactive activities which could not execute due to
unmet pre-conditions
3. Method description
This section briefly describes the method and
definitions in [13] that are utilized for the analysis and
examples in section 4. Based on the influence of
changing product parameters in a design activity on
other design activities, an impact DSM is defined.
Then it is converted to Probability DSM; and finally
reordered [11]. Unlike the typical DSM, diagonal
elements that indicate self-iteration probability, (i.e.,
the probability that activity should be reworked again)
are estimated and added. Self-iteration imply that the
work done was not approved, e.g., some aspect of the
design was not approved by the reviewer, or a piece of
software code did not pass its unit test (e.g., in Extreme
Programming (XP) [4]).
Translating the ordered DSM into a Process was
done (Figure 2): (1) Translating the DSM into a Design
Process Matrix (DPM); adding logic activities. (2)
Converting DPM into an equivalent Current process
scheme (C- Process). (3) Using the process scheme for
creating a Run Time process (RT- process).
3.1. Validity requirements
The basic validity requirements presented were:
(1) The Product Design Process has the
characteristics of a project, thus the process should
have a defined start and defined end.
(2) From any given process status, the process
should be able to reach the process end (i.e., reaching
End activity which implies process end status).
(3) When the process reached the termination state
(End activity executes), this should imply:
a. all the Design Activities (and all their
iterations) have completed; and
b. each activity has been performed at least
once.
(4) The process should be traceable.
(5) Despite the iterative nature of the process, which
enables infinite loops, the process should be enforced
to complete in a finite time.
3.2. Implementation Rules:
The implementation rules may have sub options that
are defined as Business Rules (BR), i.e., they should be
chosen according to the business environment and
business constraints considerations (e.g., Time versus
Resources). A design activity X has duration property
and it may iterate. The activity duration, duration(X), is
not described by the DSM matrix. The self-iteration
probability is p=Px, on the DSM diagonal. Translating
a single activity DSM to a DPM implies adding the
logic activities: Begin (B), End (E), Input logic (IL),
and Output logic (OL).
The following Implementation Rules (IR) apply:
(IR 1): Logic activity duration is zero; it does not have
self-iterations.
(IR 2): Design activity has one forward input links
(from IL) and one forward output link (to OL).
(IR 3): Logic activities may have multiple input or
multiple output links, but at least one forward
input link and one forward output link.
The case of one activity is depicted in Figure 2. The
initial one activity DSM (a) is translated to DPM (b).
Logic activities are added: B, E, IL, and OL. The
marking 1 in the DPM represents a link with
probability p=1. The self-iteration feedback probability
p=P, indicates the iteration of X, and is set between the
Output logic activity and the Input logic activity.
The logic applied in the IL is Join–Or, i.e., signal
from B or a feedback signal from the OL. Respectively,
the logic applied to the OL (decision procedure) is
Split-Xor; either sending feedback or a signal to the E.
The resulting Current process scheme, Figure 2(c),
is directly inferred from the DPM. Each off-diagonal
element becomes a link. Input logic (i) and Output
logic (o) are explicitly indicated, (they will be
implicitly assumed in following figures). Due to the
possible iterations, the Run Time process in Figure
2(d) has potentially infinite number of configurations,
according to the number of repetitions. In [13], the
number of iterations is practically limited by setting a
threshold on the feedback probability. Setting P
min
as
threshold derives the maximal number of iterations,
where ceil function is the nearest greater integer.
max min
( og( ) log( ))N ceil l P P=
.
Two activities may be parallel independent, serial or
coupled. The parallel independent case induced the
following IRs, being introduced as a procedure that
handles multiple parallel initiation and parallel
termination of multiple paths simulation [13].
(IR 4): The logic of Begin activity is Split- And.
(IR 5): The logic of End activity is Join-And.
(IR 6): If an activity has no previous source (in link), it
should be linked from the Begin activity.
(IR 7): If an activity has no target (out link), it should
be linked to the End activity.
The following symbols were defined [12]: Logic
indication: Split (
) for Output logic; Join () for
Input logic. Logic operations: + (OR);
(AND);
(XOR, Exclusive-Or).
The short form of multi-variable logic operations:
Multiple-Or
()
12in
A
AA A=+++
"
; Multiple-And
(
)
12in
AAA A=••
"
; and Multiple-Xor
(
)
12in
A
AA A⊗="
, where A
i
represents a link
signal, which can be a forward link F
i
or Iteration
(feedback) link I
i.
. Using the above symbols, the
following formulation was presented for the
implementation of Begin and End:
Begin
(
)
i
F
(1)
End
(
)
i
F
(2)
Forward links and feedback (iteration) links may
have distinct logic operands; the following basic rules
were defined for both IL and OL [13].
(IR 8): Forward links to design activities (lower part of
the matrix) have AND logic, on first iteration.
(IR 9): Forward link to the End activity, have XOR
(Exclusive OR) logic with the other links.
(IR 10): Feedback (iteration) links have OR logic.
In simple cases, the Input logic defines accepting
signal from all forward links (Join-And), i.e., the
activity starts once all its precedent activities have
completed; or a signal from any of the feedback links is
available; or both, (cf. Eq. 3). The Out logic procedure
may send signals to any feedback link, or (exclusively)
send signals to all forward links, but not both (cf. Eq.
4).
(
)
(
)
(
)
(
)
ii
IL F I⇐+
(3)
(
)
(
)
(
)
(
)
ii
OL I F⇒⊕
(4)
X
X P
(a) DSM (c) Current Process
(b) DPM (d) Run Time process
Figure 2. Full process description
3.2.1. Serial activities. The validity requirement 3b (in
Section 3), indicating one execution at least, imposes
different logic requirements on the first execution of
the design activity versus further iterations. For
example on its first execution, the activity cannot send
forward signal to End activity; sending a forward
signal to the next serial activity is required, otherwise
the next activity will not be performed. Sending signals
to both hinders IR 9, and was shown to cause invalid
processes.
Other implementation issues have several options
that are defined by Business Rules (BRs) numbered as
sub cases of the IRs (see appendix A for relation of
BRs and logic formulation equations).
(IR 11): OL options: sending completion signal to
following serial design activities by:
(BR 11.1): Sending only once the activity has
completed all its iterations; or
(BR 11.2): Sending once the activity has completed its
first iteration (i.e., early start of next activity).
Yet, sending a completion signal to E can be
done only once all its iterations have completed
(i.e., IR 9).
(IR 12): OL: signal to E. Second (or later) iteration of
an activity (e.g., X), may or may not send signal
to E, while the following serial activities (e.g.,
Y) have started execution (or have completed):
(BR 12.1): On second (or later) iteration, the activity
must be followed by its next serial activities.
(BR 12.2): On second (or later) iteration of the activity,
the next activities may follow, or the E may
follow (not both, subject to IR 9).
There are 4 BR combinations that change the OL,
and are manifested in the DPM. The first was the case
of BR 11.1 and BR 12.1, which was already described
by Eq. (4). The combination BR 11.1 and BR 12.2 is
formulated in Eq. (5); BR 11.2 and BR 12.1 in Eq. (6);
and BR 11.2 with BR 12.1 in Eq. (7).
OL (Σ (Ii) Σ (Fi)) End
(5)
OL (Σ(Ii) + Π(Fi)) End
(6)
OL (Σ(Ii) + Σ Fi)) End
(7)
Allowing early start option (BR 11.2) derives Run
Time options (IR 13) not described by the DPM.
(IR 13): Iterations of same activity cannot co-occur:
(BR 13.1): While the activity is executing, link signals
are directed to this activity (i.e., to its prior IL).
(BR 13.2): While the activity is executing, link signals
are directed to next iteration of the activity
(which cannot start until current iteration ends).
In the context of administrative processes, such
option was defined as Lack of Synchronization conflict
[17]. The first option (BR 13.1) entails defining the
consequences of the additional iteration rework.
Typically, the activity duration will increase.
Recalculating the activity duration in this case is
similar to the overlapping activities case [7], and can
be addressed by the same approaches.
3.2.2. Coupled activities. Coupled activities may have
serialization requirements (e.g., testing should follow
design activities). Starting coupled activities in parallel
result in processes whose OL is similar to the case of
BR 11.2 (parallelization of the iterations of serial
activities); but with different IL.
(IR 14): Coupled activity execution start:
(BR 14.1) (serialization): coupled activity may start,
after its previous activity (according to DSM)
has completed at least once.
(BR 14.2) (parallel): coupled activity may start in
parallel to all other activities in the same loop.
The Input logic for BR 14.2 is formulated in Eq. (8),
(replacing Eq. (3)); and is equivalent to the parallel
process case [20].
IL Begin + Σ(Ii) + Σ(Fi) in loop
+ Π (other forward links)
(8)
The options described in IR 12, are respectively
replaced by the options of IR 15.
(IR 15): Sending signal on completion of a coupled
design activity is done according to one of the
following business rules:
(BR 15.1): A coupled design activity should link to the
next activity on its completion.
(BR 15.2): A coupled design activity may link to the
End activity on its completion.
4. Examples and results
A comprehensive interpretation of the DSM-based
plan is given for the case of two design activities. First,
an enumeration approach is used to generate potential
processes, and filter those that comply with the
validation rules. Then, examples of process logic
problem are depicted, justifying some of the rules.
4.1. Validating the case of two design activities
Applying at most one iteration (maximum 2
executions), results in 74 distinct valid Run time
processes, which exhibit some non-trivial behavioral
aspects. The validation analysis is done using the
implementation rules applied to all potential 128
process combinations.
4.1.1. Run Time cases enumeration. For maximal
number of iteration N
max
=0, we have exactly one
execution of each design activity (as enforces by
validation rule 3(b)). The number of activities in this
case is exactly four (Begin, X, Y, End). Prcess
generation options for building a Run Time process
with two design activities are summarized in Figure 3.
Link
Option
Process
scheme
Description Applicable
rules/ Eq.
BX B-X Default link to first
activity from B
IR 6 + BR
14.1 / Eq. (3)
Bs
Split: from Begin to
multiple activities in
parallel
IR 6 + BR
14.2 / Eq. (8)
XY X–Y Final iteration of an
activity links to next
activity
BR 12.1 or BR
15.1 / Eqs. (4),
(6)
XE X-E Final iteration of an
activity can link
directly to E
BR 12.2 or BR
15.2 / Eqs. (5),
(7)
YX Y–X Final iteration of an
activity which can
feedback to other
activity in loop
BR 14.2 / Eqs.
(6), (7), (8)
YE Y-E Final iteration of an
activity without target
should link to E
IR 7 / Eqs. (4),
(5), (6), (7)
Figure 3. Optional run time linkages
There are two options for starting the process, link
from B to X (BX), or split to both design activities
(Bs); two options to link X after its completion: link to
E (XE), or link to Y (XY); and similar options for Y (YE,
YX). The options presented for X and Y are the only
applicable options of the final iteration of an activity
(i.e., it can not iterate to itself in this case, N
max
=0).
Some options are applicable only for specific business
rules and applicable equations, as indicated.
There are eight potential Run Time processes,
utilizing the three choices: Begin (2 options), X (2
options), Y (2 options). However, validity requirements
filter out some of them. The examples are depicted in
Figure 4. For easier display, we assume the Input and
Output logic are embedded in the activity (short RT
process description).
The option (BX+XE) is invalid since Y activity is
not executed. Using ‘*’ as don’t care symbol, we can
defined the choices as {BX, XE, ‘*’}, i.e., the process
is equivalent for both choices of Y option. In general,
any serial process combination, which includes only
iterations of activity X is invalid, Figure 4(a).
(a) Invalid: No Y activity (d) Valid: Coupled
(b) Valid: Serial (e) Invalid: (X-Y) + (X-E)
(c) Invalid: (X-Y) + (Y-X),
no End
(f) Valid: Coupled, BR
13.1
Figure 4. Run Time (short), No iterations
The serial valid case is depicted in Figure 4(b). The
combination of options (XY+YX), Figure 4(c), is
always invalid since the process does not reach the End
activity, violating validity requirement (1). This
combination occurs for two cases of process start {‘*’,
XY, YX}.
Valid Parallel options are depicted in Figure 4(d)
and (f). In the latter case there should be a difference in
the activities duration, applying rule BR 13.1 (rule BR
13.2 cannot apply since there are no additional
iterations). If both activities have exactly the same
duration, the options in Figure 4(f) are reduced to the
case in Figure 4(d). The Run Time process cases
described above are all the eight potential cases derived
from the options in Figure 3.
An additional invalid option, which might
graphically look like other Run Time processes, is
depicted in Figure 4(e). Such option cannot be
generated from the options presented in Figure 3; X
cannot link to both Y and End. Using business rules
interpretation, proceeding to End activity cannot be
done in parallel to proceeding or feedback to other
activities (Xor logic). Proceeding to End should mean
completion, while feedback or continuing to other
activities has a contradicting intent.
4.1.2. Run Time with iterations. Allowing one
iteration of each activity, N
max
=1, either self-iteration
or iteration due to the other activity, yields maximum
six activities: Begin, first X iteration, first Y iteration,
last iterations of X and Y, End. In addition to the
linkage options described in Figure 3, the interim
iterations (i.e. first activity execution) have total of four
options detailed in Figure 5.
Link
Optio
n
Process
scheme
Description Applicable
rules/equations
XX X-X Self-iteration Eqs. (4), (5),
(6), (7)
YY Y–Y Self-iteration Eqs. (4), (5),
(6), (7)
Xs
Split: Iteration and
forward link, using the
applicable business
rules
BR 11.2 or BR
14.2
Eqs. (4), (6)
Ys
Split: Choosing multiple
iterations (feedback
links have OR logic)
Eqs. (4), (5),
(6), (7)
Figure 5. Additional linkage options for iteration
There are five choices to make based on the linkage
options detailed above: Begin activity (two options),
first activity iteration X and Y (four options each), and
final activity iteration X and Y (two options each). The
combinations yield 128 potential processes, which
include invalid processes.
Many potential Run Time processes are invalid due
to the combination (XY + YX). This choice, indicated
by {‘*’, ‘*’, ‘*’ XY, YX}, appears in quarter of the
cases. However, the process may end, and might be a
valid process, before reaching these choices, (e.g. {Bs,
XE, YE, ‘*’, ‘*’}), or might be invalid before reaching
these choices, thus the actual number of invalid
processes due to this choice is sixteen. Further sixteen
invalid cases are described by the process path B-X-E,
using the selection {BX, XE, ‘*’, ‘*’, ‘*’}. These
sixteen cases include four cases of (XY + YX), and are
counted as part of the B-X-E combination since the
process does not reach the choices (XY + YX). Eight
invalid options, with 3 executions of X are includes in
the process path B-X-X-E: {BX, XX, ‘*’, XE, ‘*’}.
Additional four invalid cases are depicted in Figure
6. Two cases refer to the selection {BX, XX, XY, YX,
‘*’}. These two cases are equivalent and yield an
invalid process in Figure 6(a). Two additional cases
refer to selection {BX, XX, XY, Ys, ‘*’}, in which after
X has already iterated, linkages from Y are connected to
both X and Y, thus a third iteration of X is required. The
case {BX, XX XY, Ys, YE} could complete as a valid
process, Figure 6(b); the case {BX, XX, XY, Ys, YX}, is
invalid as discussed above.
Besides invalid options, there are distinct options
which result in the same valid Run Time process, thus
further reducing the count of distinct valid Run Time
processes. For example, the selection {BX, XY, YE, ‘*’,
‘*’}, representing four options, is equivalent to one
valid process depicted in Figure 4(b). Selecting {Bs,
XE, YE, ‘*’, ‘*’}, produce the process in Figure 4(d).
(a) Invalid process (b) Invalid: 3
rd
X iteration
Figure 6. Invalid: 3
rd
X iteration
Applying business rules BR13.1 or BR 13.2 may
add distinct Run Time results. For example, if both
iterations of Y exist, then the link XY may result in X-
Y(1) or X-Y(2), where Y(i) indicate the iteration
number.
The total count is 128 potential Run Time cases,
minus (16+16+8+4) invalid options, yield 84 valid
cases. Out of these there are 28 repeated cases (i.e.,
equivalent to others). The application of business rules
BR13.1 or BR 13.2 adds 18 distinct run time cases. In
total there are 74 different Run Time processes, which
are the result of applying different combinations of
business rules to a DSM with two design activities and
maximum of two executions for each activity.
4.2. Logic verification issues
In [13], examples were used to demonstrate the
implications of the above mentioned business rules for
simulating parallel activities. The utilization of
implementation rules (IR 4 to IR 7) was discussed in
the context of parallel start of multiple activities and
parallel completion. The simulation problems arising if
these rules are not applied were presented.
The examples in Figure 7 illustrate the
implications of adding self-iterations. A short version
of the DPM is used: Begin logic activity (B), and End
activity (E) are explicitly presented; Input logic and
Output logic activities are assumed implicitly before
and after the design activities X and Y, respectively.
In Figure 7(a), X and Y activities have self-
iteration linkages. According to Eq. (3), the Input logic
of X is IL
B+X, i.e., signal from Begin or iteration
(from Out logic activity). In a similar manner, the input
logic of Y, is IL
X+Y. The Output logic of Y is Split-
Xor between iteration and the link to End activity (IR
9); other options are invalid. There are two Output
logic options for X: either all iterations should
complete before continuing to Y (BR 11.1), or
(parallelization case) additional X iterations may
execute in parallel to Y (BR 11.2). In the second case,
the logic of E requires that the last X iteration will
complete, though no link exists from X to E in the
DPM. This is required since Y may complete and send
signal to E, but a new iteration of Y may later start due
to rework from X. In general, requiring completion of
all activity iterations is derived from validation
requirement 3(a).
In Figure 7(b), the option to “jump” to the End
activity on the second (or following) iteration of X was
added, i.e. applying BR 12.2. Similar to the previous
case, the second iterations X
2
(or following) has two
options (BR 11.1 and BR 11.2). E has to wait for X to
complete all its iterations as in the previous case.
Cas
e
DPM Valid Logic Invalid Logic
examples
(a)
BXYE
B0
X1 0.1
Y 1 0.1
E 1 0
X
X
Y;
(BR 11.1)
X
X+Y;
(BR 11.2)
Y
Y
E;(IR 9)
E
X
Y
Y
Y+E ;
Y
Y
E ;
(b)
BX Y E
B0
X1 0.1
Y 1 0.1
E
1
1 0
(BR 12 .2):
X
2
X
Y
E ;
X
2
(X+Y)
E;
X
2
(X
Y)+E ;
X
2
X+Y+E
;
And more
options
(c)
BX Y E
B0
X 1 0.1 0.1
Y 1 0.1
E 1 0
X
B+Y ;
Y
X
Y
E ;
Y
(X+Y)
E
X
B
Y ;
Y
X+Y+E ;
Y
X
Y
E ;
Figure 7. Additional linkage options for iteration
Figure 7(c) represents a coupled pair of activities.
The input logic of X cannot be Join-And since this will
cause a deadlock (X waits for Y, while Y waits for X).
The valid output logic options of Y are similar to the
second iteration of X
2
in the previous case. The identity
of logic rules being applied for feedback (from Y) and
for the second iteration (of X) was presented in [6].
However, it reflected only second iteration of X in a
coupled case and not due to self-iterations.
5. Discussion and Conclusions
Since self-iterations are not addressed in DSM-
based simulation literature, the actual simulations have
a reduced scope of potential run time processes. The
richness of run time processes due to self-iterations
was demonstrated, establishing the multiple variations
of interpreting the DSM linkages to current process
plans. The example illustrates the complexity of
enumerating the run time cases, even for a very simple
case; and demonstrates the complexity of verifications
of iterative processes.
Verifying rules for the case of two activities and
demonstrating their applicability, do not imply other
cases correctness. However, such exercise does
elucidate the rudiments of iterative process validation,
demonstrating the ability to constructively build correct
processes, and directs further research endeavor.
.
6. Appendix A
The relations between IRs, BRs, and the IL and OL
equations is shown in Figure 8. IL is described by Eqs.
3 and 8. The OL of the various cases is formulated in
Eqs. 4 to 7.
Figure 8. Relations between implementation and
business rules
7. References
[1] Aalst, W.M.P. van der, The Application of Petri Nets to
Workflow Management, The Journal of Circuits,
Systems and Computers, 8(1):21–66, 1998.
[2] Aalst W.M.P. van der, Hee K.M. van, Workflow
Management: Models, Methods, and Systems, MIT
Press, Cambridge, MA. 2002.
[3] Aalst W.M.P. van der, Hofstede A.H.M. ter,
Kiepuszewski B., Barros A.P., Workflow patterns,
Distributed and Parallel Databases 14(1):5–51, 2003.
[4] Beck K., Extreme Programming Explained, Boston:
Addison-Wesley, 2000.
[5] Browning T.R., Applying the Design Structure Matrix
system to decomposition and integration problems: A
review and new directions, IEEE Trans. Eng. Manage.,
48:292-306, 2001.
[6] Browning T.R., Eppinger S.D., Modeling impacts of
process architecture on cost and schedule risk in product
development, IEEE Trans. Eng. Manage., 49(4):428-
442, 2002.
[7] Cho S.H., Eppinger S.D., A simulation-based process
model for managing complex design projects, IEEE
Trans. Eng. Manage., 52(3): 316-328, 2005.
[8] Clarkson P.J., Melo A.F., Eckert C.M., Visualization of
Routes in Design Process Planning, In Proceedings of
the Fourth International Conference on Information
Visualisation (IV2000), IEEE Computer Society, pp.
155-164, London, 2000.
[9] Eppinger SD, Whitney DE, Smith R, Gebala D, A
model-based method for organizing tasks in product
development, Res. Eng. Design, 6(1):1-13, 1994.
[10] Eppinger SD, Nukala MV, Whitney DE, Generalized
models of design iterations using signal flow graph, Res.
Eng. Design, 9:112-123, 1997.
[11] Karniel A, Belsky Y, Reich Y, Decomposing the
problem of constrained surface fitting in reverse
engineering, Computer-Aided Design, 37:399-417,
2005.
[12] Karniel A., Reich Y., From DSM based planning to
Design Process Simulation: A review of process scheme
verification issues, submitted, 2006(a)
[13] Karniel A., Reich Y., A Coherent interpretation of
DSM plan for PDP simulation, submitted, 2006(b)
[14] Lévárdy V., Browning T.R., Adaptive test process –
designing a project plan that adapts to the state of a
project, Fifteenth Annual International Symposium of
the International Council on Systems Engineering
(INCOSE), 2005.
[15] Reising W., Rozenberg G., editors, Lectures on Petri
Nets I: Basic Models, Lecture notes in Computer
Science vol 1491, 1998.
[16] Rinderle S., Reichert, M., Dadam, P., Correctness
Criteria for Dynamic Changes in Workflow Systems - A
Survey, Data and Knowledge Engineering, 50(1):9-34,
2004.
[17] Sadiq W., Orlowska M. E., Applying Graph Reduction
Techniques for Identifying Structural Conflicts in
Process Models, Lecture Notes in Computer Science,
1626:195-209, 1999.
[18] Sadiq S., Orlowska M.E., Sadiq W., Foulger C., Data
flow and validation in workflow modeling, Proceedings
of the International Conference in Research and
Practice in Information Technology, ADC’2004,
Dunedin, New Zealand, 2004.
[19] Sered Y., Reich Y., Standardization and modularization
driven by minimizing overall process effort, Computer-
Aided Design, 38(5):405-416, 2006.
[20] Smith R.P., Eppinger S.D., Identifying controlling
features of engineering design iteration, Management
Science, 43(3):276-293, 1997 (a).
[21] Smith R.P., Eppinger S.D., A predictive model of
sequential iteration in engineering design, Management
Science, 43(8):1104-1120, 1997(b).
[22] Steward D.V., Systems Analysis and Management:
Structure, Strategy, and Design, Petrocelli Books, NJ,
1981.
[23] Yassine, A., Investigating product development process
reliability and robustness using simulation, Journal of
Engineering Design, in press, 2006.
[24] Yassine A., Braha D., Complex concurrent engineering
and the design structure matrix method, Concurrent
Engineering Research and Applications, 11(3):165-176,
2003.
... A DSM net has Design Activities (representing the activity of designing a product item), and Logic Activities (that represent the implementation of process logic). Design activities may have self-iterations (e.g., the checker did not approve the design, or it failed in testing), which are typically not addressed in the DSM literature and were presented in [16]. The DSM net (that might change in time) represents the Current process scheme, i.e., the potential options that are applicable and aligned with the current process state. ...
... Self-iteration rules apply to the dynamics of the process, i.e., once activity iteration occurs [16]. Applying the rule option do not change the DSM net, but shapes the runtime process. ...
Conference Paper
Full-text available
Planning a New Product Development (NPD) process has an evolving nature. It is repeatedly updated during the development due to changes in requirements, technology, product concept, or testing results that drive product design modification and require changes in the plan. The Design Structure Matrix (DSM) is utilized to generate a process plan that is based on the product knowledge. The translation of the DSM-based plan to a DSM net process scheme requires implementation rules. Such translation is not unique and there are implementation choices we define as business rules. This paper presents the implementation rules used for dynamic changes in the plan. The application of these rules conforms to a correctness criteria based on the soundness criteria used in Petri nets for process scheme verification.
... Resource scheduling, based on DSM plan was applied to assignment of computer resources in real time calculations [26]; and to a weekly assignment of project resources in [27]. The typical DSMbased planning algorithms do not utilize the diagonal [28]. However, if there is only one simulation parameter involved per each activity, the diagonal cells could be used to present it. ...
Conference Paper
Full-text available
Simulations of New Product Development (NPD) processes using their specific contexts can provide project managers with decisions-making aids. The NPD context, which incorporates knowledge about the product, requirements, technology, and other factors, is dynamically evolving during the process. The Design Structure Matrix (DSM) could be used to model this product knowledge and reordering algorithms could be used for process planning. The plan should be updated as the product knowledge evolves; however, the transition from DSM-based plan to process scheme implementation is not unique. The process can then be used for simulating process related measures that can guide the decision making regarding the preferable implementation strategy. However, such decisions should reflect the statistical confidence of the simulation results. In the current work, the properties of multiple repetitive simulations are used for supporting decision-making based on statistical confidence interval derived by evaluating the significance of the difference between the results of applying different DSM-based plan implementation options (defined as business rules). The same approach is applicable to similar decision-making related to processes in general.
... If BR 15.1 was used, the latter link was not included and the resulting C-process would be as in (c). Additional implications of utilizing self-iterations were further studied in [19] and provided additional insight to iterative processes modelling. ...
Conference Paper
Full-text available
Design Processes are highly complex and fundamentally iterative. The increasing knowledge about the product, while designing, manifests in design changes of previously accomplished activities. Such iterations are considered as a major source of increased product development lead-time and cost. Planning the Design Process is challenging, and requires simulation tools. The Design Structure Matrix (DSM) is a known method for process planning. Since the DSM itself does not express all the required information for defining the process logic, process logic interpretation is required. " Business Rules " , being logic interpretation options that are applicable in different business cases, can guide automatic translation of the DSM to valid iterative process plans. There is a gap in the literature between DSM-based process planning, and process modelling literature concerning the verification of processes. A consistent method of transformation from a DSM model to a logically correct, concurrent process model, in the case of iterative activities, is lacking. Such consistent interpretation of DSM plan to a valid process scheme is proposed in this study. The applicable logic interpretation options are expressed by " Business Rules " that convey different business cases, and can guide automatic translation of the DSM. The proposed framework subsumes logic interpretation types found in the DSM literature.
Chapter
The following section presents the proposed process for gathering DSM data. As in other DSM-based process planning methods, the product data is deriving the process plan. A key issue in the current work is the utilization of data changes throughout the design process. The detailing level is not pre-defined and is changing and adapting to the applicable product knowledge. Collecting the data throughout the design process requires a tool. An existing applicable tool could be a PLM product, where other product relation links (e.g., father-son) are managed. A specific tool for collecting the data may also apply but would be less effective. The process of collecting the DSM data as part of the overall DnPDP is fully described in Sect. 8. 5. The following example reflects the initial knowledge at the conceptual design stage.
Chapter
This chapter describes the concepts, methods, and enhancements that compose this research. Due to the variety of subjects, simple examples are added to each issue, as well as simple integrated examples. The implementation of a fully integrated example is presented in Chap. 1. Section 8.1 presents the main DnPDPDynamic new-Product Design Process (DnPDP) frameworkDnPDP framework concepts required for managing dynamic process changes of NPDNew Product Development (NPD) processes. The proposed integrated framework (Karniel and Reich 2007b) is presented as a closed-loop process (Sect. 8.2), extended to a decision-making process by additional control loop (Sect. 8.3). The computational view of the framework separates it into process steps forming the controller Meta-process (Sect. 8.4). The controller accepts product data as feedback of the design process. The (changing) product data is converted to a process scheme. The main conversion steps go from product data to Probability DSM,DSM:Probability then to Design Process Matrix (DPM)Design Process Matrix (DPM), and then to process scheme. These steps, defined as the planning cycle, are overviewed in Sect. 8.5. The evolving process models are discussed in Sect. 8.6.
Chapter
The main applications presented in the following example are the monitoring of a dynamically changing process scheme; the implication of the simulation of dynamic planning capabilities; and the potential use of the simulation results for process-related decision-making.The example represents an NPD environment, where process activities are derived from the product structure and the relations between product components, which are subject to changes during the design as new knowledge becomes available.
Chapter
The interpretation of the DSM to a process scheme is not unique, i.e., the presented links between activities require additional interpretation for translating the DSM into a logically-correct executable process that can be simulated. Implementation Rules are used for interpretation. The Implementation Rules (IR) may have suboptions defined as Business Rules (BR), i.e., they should be chosen according to the business environment and business constraints considerations (e.g., time versus resources).
Chapter
The following sections describe the integration of the DSM planning model with process modeling approaches of Petri netsPetri net. First, the process correctness criteriaCorrectness criteria for the Dynamic new-Product Design Process (DnPDP)Dynamic new-Product Design Process (DnPDP) are presented in Sect. 9.2. Most requirements are adapted from WF-netWF-nets correctness criteriaCorrectness criteria.
Chapter
Complex and knowledge intensive innovation, engineering and service projects are important for companies to improve their market performance. Project planning and scheduling, resource assignment and cost planning set pointers for a successful project implementation. Due to complex interdependencies between project activities and resources, increasingly less project managers are able to deal with the complex and latent uncertainties of the planning process. Discrete event, Monte Carlo simulation seems to be a promising approach to comply with the requirements of realistic planning and scheduling of complex projects. In this paper an activity-oriented, personnel-integrated simulation model is presented facilitating project managers to model and simulate the schedule of weakly-structured complex projects like engineering and service projects and, thus, evaluate the strengths and weaknesses as well as opportunities and risks of these projects. Two explorative industry case studies in the chemical industry and metal processing industry are presented to demonstrate successful operational application of the simulation-based project planning and scheduling approach and the potential of respective project risk assessment.
Conference Paper
Full-text available
A Product Development Process (PDP) is a collection of interconnected activities targeted to convert a new idea, concept, or market opportunity, into a marketed product. PDPs are highly complex, dynamic, and iterative. They are complex because they involve multiple disciplines contributing to the development of multidisciplinary products (e.g., machatronics), with limited resources, shortened development time, and increased quality and regulatory demands. They are dynamic, since they evolve continually due to various reasons, including market changes, technological changes, and organizational changes [10]. PDPs are inherently iterative, as changes in the design of one component may result in changing the design of other components; e.g., testing activities set a-priori planned iterations [7]; and unplanned changes or problems add unpredicted feedback iterative rework [10]. Iterations are considered as major source of increased product development lead-time and cost [1]. Amongst all PDPs, those related to new products, i.e., New Product Development (NPD) processes present additional complication to product developers: developers' knowledge of the new product being designed is initially poor and only increases during the development process. Given this, we cannot hope to define the scope of work needed for a new product, nor could we plan how to manage it a-priori [10]. Therefore, change management and its derivative, design iterations, become major concern in NPD processes and in particular, in product design activities [11]. Planning the process and performing project activities in an appropriate sequence is critical in order to minimize rework due to information interdependencies [8]. Introduced by Steward [9], the Design Structure Matrix (DSM) is utilized to capture the required product knowledge for process planning. The presented work describes the development of a framework for managing dynamic process planning for design processes called Dynamic Product Design Process (DPDP). The proposed framework supports a cyclic process from planning a design process to its execution; and re-planning based on product knowledge evolution during the design (i.e., product structure, product parametric dependencies and constraints); customer requirements, and ad-hoc changes (see Figure 1). Planning the dynamic process scheme is derived from the DSM method. The product knowledge (product structure and design relations between components) is transferred to design activities by assigning a design activity to each product component. A relation value indicates an estimation of the consequence of making a design change on the need to make design changes in other components. These estimations are converted to probabilities [12], yielding Probability DSM, and the activities are reordered [2] to create a process plan, still in DSM representation.
Conference Paper
Full-text available
Design Processes are highly complex and fundamentally iterative. The increasing knowledge about the product, while designing, manifests in design changes of previously accomplished activities. Such iterations are considered as a major source of increased product development lead-time and cost. Planning the Design Process is challenging, and requires simulation tools. The Design Structure Matrix (DSM) is a known method for process planning. Since the DSM itself does not express all the required information for defining the process logic, process logic interpretation is required. " Business Rules " , being logic interpretation options that are applicable in different business cases, can guide automatic translation of the DSM to valid iterative process plans. There is a gap in the literature between DSM-based process planning, and process modelling literature concerning the verification of processes. A consistent method of transformation from a DSM model to a logically correct, concurrent process model, in the case of iterative activities, is lacking. Such consistent interpretation of DSM plan to a valid process scheme is proposed in this study. The applicable logic interpretation options are expressed by " Business Rules " that convey different business cases, and can guide automatic translation of the DSM. The proposed framework subsumes logic interpretation types found in the DSM literature.
Article
Full-text available
This paper uses the design structure matrix to model and simulate the performance of product development processes. Although simulation is a powerful tool for analysing process performance, its ability is limited by the availability of accurate input information. Design structure matrix simulation requires process data that is hard to assess or estimate directly from development participants. In this paper, we propose a methodology that allows a more practical estimation of an important simulation input parameter: rework probability. Furthermore, we show how this assessment method (combined with simulation) allows managers to evaluate process improvement plans based on two resultant process measures: reliability and robustness. The method is illustrated with a real application from the automotive industry.
Conference Paper
Order cycle time is an important issue in supply chain management and Hard to control. It is a set of either deterministic or probabilistic activities. Since order cycle is assumed as a GERT network, has a variable time and a branching probability for each arrow, hence the total cycle time is variable. This paper proposes a new solution technique based on GERT method to control Customer lead-time in a supply chain The proposed method is efficient in estimating mean and variance of the order Cycle time and more accurately confidence interval.
Conference Paper
In this paper we discuss the application of the GERT (Graphical Evaluation and Review Technique) network to business processes. Since the introduction of GERT, many papers have been published to show its application in various fields of engineering, management and system study. Application to information systems in general and business processes in particular has not yet been shown in the literature. One of the strengths of the GERT networks is the graphical representation, which is intuitive and easy to understand. The GERT network model, proposed in this paper, is adapted for business processes and information systems modeling, which means extending the graphical representation of GERT networks with respect to the peculiarities of business processes. We propose a new approach for the development of mathematical models of business processes using Z-transform and Mason’s rule. We apply DEMO (Dynamic Essential Modeling of Organization) to produce business process models, which are then further analyzed using GERT. As an application to real world, we illustrate our method taking the case of the work of the Conciliation Board for Consumers in the Netherlands.
Article
Conventional techniques for project planning make the flawed assumption that all activities and relations are known and can be planned a priori. Iteration models show that iteration is a fundamental aspect of novel projects, but they also assume all activities and interactions are known. Of course, some kind of hypothesized project baseline (plan) is required for management. Such plans are acknowledged to be imperfect, so management reserves are kept for schedule and budget where possible. How big should these reserves be? This paper describes research on an improved approach to process planning and control whereby activity modes are allowed to vary and dynamically recombine based on the evolving state of the project as it progresses. If the project is behind schedule or over budget, then activities or iterations which would seem prudent from a technical performance standpoint may not be feasible as originally planned. Similarly, if a project is ahead of schedule or under budget, additional improvements versus the original plan may add value. What is prudent depends on the effect of any action on the overall value of the project. As such decisions are made under uncertainty, the proposed approach accounts for uncertainties and their impacts (risks) in the decision. By assuming from the outset—while formulating the project plan—that the project will adapt, it is possible to build an improved project plan, one that is more accurate, realistic and flexible. Additionally, the proposed method supports management decisions about potential corrective actions during project execution. This paper applies the approach in the context of system verification and validation (V&V) processes.
Article
This paper presents a model describing sequential iteration, one of the fundamental solution processes experienced in complex engineering design projects. The model is based upon the design structure matrix representation and assumes that each individual design activity is of deterministic duration with probabilistic repetition, where the repeat probabilities are defined by the strength of the task coupling. Using the model, we are able to compute the expected duration of the iterative solution process, and to suggest an initial ordering of the coupled design tasks to minimize the expected duration. We conclude the paper with a discussion of limitations and several extensions to the sequential iteration model.