Product Family Design and Platform-Based Product D

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/226657508

Product family design and platform-based product development: A state-of-


the-art review

Article  in  Journal of Intelligent Manufacturing · February 2007


DOI: 10.1007/s10845-007-0003-2

CITATIONS READS

642 10,900

3 authors, including:

Timothy W. Simpson Zahed Siddique


Pennsylvania State University University of Oklahoma
398 PUBLICATIONS   16,934 CITATIONS    180 PUBLICATIONS   1,995 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Investigating the Effectiveness of Machine Learning Paradigms for Supporting Engineering Designers in Rapidly Evolving Digital Manufacturing View project

National Science Foundation Additive Manufacturing Education View project

All content following this page was uploaded by Zahed Siddique on 13 February 2015.

The user has requested enhancement of the downloaded file.


J Intell Manuf (2007) 18:5–29
DOI 10.1007/s10845-007-0003-2

Product family design and platform-based product development:


a state-of-the-art review
Jianxin (Roger) Jiao · Timothy W. Simpson ·
Zahed Siddique

Received: August 2005 / Accepted: February 2006 / Published online: July 2007
© Springer Science+Business Media, LLC 2007

Abstract Product family design and platform-based To compete in the marketplace, manufacturers have
product development has received much attention over been seeking for expansion of their product lines and
the last decade. This paper provides a comprehensive differentiation of their product offerings with the intu-
review of the state-of-the-art research in this field. A itively-appealing belief that large product variety may
decision framework is introduced to reveal a holistic stimulate sales and generate additional revenue (Ho &
view of product family design and platform-based prod- Tang, 1998). Initially variety does improve sales as the
uct development, encompassing both front-end and offerings become more attractive; but as variety keeps
back-end issues. The review is organized according to increasing, the law of diminishing returns suggests that
various topics in relation to product families, includ- the benefits do not keep pace (Wortmann, Muntslag, &
ing fundamental issues and definitions, product portfolio Timmermans, 1997). Facing such a dilemma, a company
and product family positioning, platform-based product must optimize its external variety with respect to the
family design, manufacturing and production, as well as internal complexity resulting from product differentia-
supply chain management. Major challenges and future tion (Tseng & Jiao, 1996).
research directions are also discussed. Designing and developing product families has been
well recognized as an effective means to achieve the
Keywords Product family · Product platform · economy of scale in order to accommodate an increasing
Product architecture · Customization · Variety product variety across diverse market niches (Meyer &
modularity · Commonality · Configuration Utterback, 1993; Sundgren, 1999). In addition to lever-
aging the cost of delivering variety by reusing proven
elements in a firm’s activities and offerings, product fam-
Introduction ily design can offer a multitude of benefits including
reduction in development risks and system complex-
Manufacturing companies have devoted much atten- ity, improved ability to upgrade products, and enhanced
tion to the contemporary battlefield called mass cus- flexibility and responsiveness of manufacturing processes
tomization, which aims to satisfy individual customer (Sawhney, 1998). Many companies are investing in prod-
needs by introducing product proliferation while taking uct family development practices in order to provide
advantage of mass production efficiency (Pine, 1993). sufficient variety to the market while maintaining the
economies of scale and scope within their manufactur-
J. (Roger) Jiao (B)
ing capabilities (Robertson & Ulrich, 1998).
Nanyang Technological University, Singapore, Singapore
e-mail: jiao@pmail.ntu.edu.sg A sizeable body of research on product family design
and platform-based product development has been
T. W. Simpson reported over the last decade (Simpson, 2004). The lit-
Pennsylvania State University, University Park, PA, USA
erature involves many, if not all, aspects of product
Z. Siddique fulfillment within a manufacturing enterprise. Product
University of Oklahoma, Norman, OK, USA family design and development has been tackled from
6 J Intell Manuf (2007) 18:5–29

various perspectives as well, such as in the areas of busi- customer and functional domains constitutes the front-
ness strategy, marketing, manufacturing and production, end issues associated with developing product families.
customer engineering, information technology, and gen- Such a product family definition task is always carried
eral management. This paper provides a comprehen- out within an existing product portfolio and manifests
sive review of the state-of-the-art research regarding itself through those common practices of order con-
product family design and development. To proceed, figuration and sales force automation. Product family
a holistic view of product family development is pre- design solutions are generated in the physical domain
sented first, encompassing both front-end and back-end by mapping FRs to design parameters (DPs) based on
issues of product families. With respect to the outlined the shared product platform. This stage involves typical
decision framework of product family development, the decisions regarding product family design and configu-
literature review is organized in the subsequent sections ration. At the front-end, the product portfolio articu-
according to various topics in relation to product fam- lates detailed achievement of customer satisfaction in
ilies, which emphasize on the fundamental issues and the customer domain in the form of specifications of
definitions, product portfolio and product family posi- functionality in the functional domain. On the other
tioning, product family design, manufacturing and pro- hand, the main focus of platform-based product fam-
duction, and supply chain management, respectively. ily design is the technical feasibility of DPs in terms of
Finally key challenges and future research directions fulfilling the specified functionality.
are outlined. The back-end issues associated with product fami-
lies involve the process and logistics domains, which are
characterized by process variables (PVs) and logistics
Decision framework variables (LVs), respectively. The mapping from DPs to
PVs entails the process design task, which must gen-
Product family design and development essentially erate manufacturing and production planning within
entails a conceptual structure and overall logical orga- existing process capabilities and utilize repetitions in
nization of generating a family of products by providing tooling, setup, equipment, routings, etc. Correspond-
a generic umbrella to capture and utilize commonal- ing to a product platform, production processes can be
ity, within which each new product is instantiated and organized as a process platform in the form of stan-
extended so as to anchor future designs to a common dard routings, thus facilitating production configuration
product line structure. The rationale lies in not only for diverse product family design solutions (Jiao, Tseng,
unburdening the knowledge base from keeping variant Ma, & Zou, 2000). Since the main concern in the process
forms of the same solution, but also in modeling the domain is manufacturability and cost commitment, pro-
design process of an entire class of products that can cess design is the de facto enabler of mass production
widely variegate designs based on individually custom- efficiency.
ized requirements within a coherent framework (Jiao & To achieve mass customization, more and more com-
Tseng, 1999). panies are moving towards assemble-to-order produc-
Figure 1 illustrates the decision framework of product tion (Wortmann et al., 1997). Outsourcing has emerged
family design and development along the entire spec- as a promising strategy for leveraging resources and
trum of product realization according to the concept of capabilities worldwide while focusing on the core com-
design domains (Suh, 2001). Based on such a holistic petency. It thus becomes imperative to coordinate
view, product family design and development encom- product and process design of product families within a
passes consecutively five domains, namely the customer, supply chain framework. The logistics domain addresses
functional, physical, process and logistics domains. Prod- the supply chain related issues of product family ful-
uct family decision-making involves a series of “what- fillment, which are mainly about supply chain config-
how” mappings between these domains. uration, resource allocation, supplier management and
The customer domain is characterized by a set of cus- supply contracting. A supply platform plays an impor-
tomer needs (CNs) representing segmentation of mar- tant role in supporting supply chain design through the
kets that demand for product families and triggering mapping from the process domain to the logistics
downstream product family design mappings in a cas- domain. The main concern is to align production con-
cading manner. The CNs are first translated into func- figuration and supply chain decisions. Throughout the
tional requirements (FRs) in the functional domain, mappings across these five domains, there are some fun-
in which designers take into account engineering con- damental issues related to product family design deci-
cerns and elaborate these requirements based on avail- sions, including conceptual implications of a product
able product technologies. The mapping between the family, product platform, product architecture, product
J Intell Manuf (2007) 18:5–29 7

Fig. 1 A holistic view of


product family design and
development

variety, modularity and commonality. The following review 1998) to being industry and product specific (Ericsson
starts with these fundamentals. & Erixon, 1999; Sanderson & Uzumeri, 1995). In addi-
tion, the meaning of platform differs in scope. Some
definitions and descriptions focus mainly on the product
or artifact itself (McGrath, 1995; Meyer & Utterback,
Fundamental issues
1993), whereas others try to explore the platform con-
cept in terms of a firm’s value chain (Sawhney, 1998).
Product family
There are two streams of research prevailing in the
field of developing product platforms. One perspective
Streams of individual products generated by firms may
refers to a platform as a physical one, namely a collec-
be thought of as evolving families of products (Meyer &
tion of “elements” shared by several products. Accord-
Utterback, 1993). A product family refers to a set of sim-
ingly, the major concern is how to identify the com-
ilar products that are derived from a common platform
mon denominators for a range of products (Wilhelm,
and yet possess specific features/functionality to meet
1997; Ericsson & Erixon, 1999). The effort is geared
particular customer requirements (Meyer & Lehnerd,
towards the extraction of those common product ele-
1997). Each individual product within a product fam-
ments, features, and/or subsystems that are stable and
ily, i.e., a family member, is called a product variant or
well understood, so as to provide a basis for introducing
instance. While a product family targets a certain market
value-added differentiating features (Moore, Louviere,
segment, each product variant is developed to address a
& Verma, 1999).
specific subset of customer needs of the market segment.
Meyer and Lehnerd’s work (1997) is the represen-
All product variants share some common structures and
tative of another dominating perspective to product
product technologies, which form the platform of the
platform. They define a product platform as “a set of
product family (Erens & Verhulst, 1997).
subsystems and interfaces developed to form a com-
The interpretation of product families depends on
mon structure from which a stream of derivative prod-
different perspectives. From the marketing and sales
ucts can be efficiently developed and produced” (Meyer
perspective, the functional structure of product families
& Lehnerd, 1997,p. 39). The major issue is to exploit
exhibits a firm’s product line or product portfolio and
the shared logic and cohesive architecture underlying a
thus is characterized by various sets of functional fea-
product platform. McGrath (1995) defines product plat-
tures for different customer groups (Agard & Kusiak,
form as a collection of the common elements, especially
2004). The engineering view of product families embod-
the underlying core technology, implemented across a
ies different product technologies and the associated
range of products. Robertson and Ulrich (1998) define
manufacturability and thereby is characterized by vari-
a platform as the collection of assets that are shared by
ous design parameters, components, and assembly struc-
a set of products. The assets include components, pro-
tures (De Lit & Delchambre, 2003; Simpson, 2004).
cesses, knowledge, as well as people and relationships.
Baldwin & Clark (2000) define three aspects of a
Product platform product platform: (1) its modular architecture, (2) the
interfaces, and (3) the standards that provide design
Product platforms have been defined diversely, ranging rules to which the modules must conform. To facilitate
from being general and abstract (Robertson & Ulrich, platform-based product family development, interface
8 J Intell Manuf (2007) 18:5–29

management is reported as a distinct process of defin- While modularity deals with the mapping from func-
ing the physical interfaces between subsystems (Sund- tions to components, integrality involves standardiza-
gren, 1999). Zamirowski and Otto (1999) discern three tion and decoupling of the interfaces between com-
types of product platforms: modular platforms, scalable ponents (Ulrich and Eppinger, 1995). Robertson and
platforms, and generational platforms. A modular plat- Ulrich (1998) observe that increasing modularity with
form is used to create variants through configuration proper integrity is conducive to the management of
of existing modules (Meyer & Lehnerd, 1997). A scal- tradeoff between distinctiveness and commonality in
able platform facilitates the differentiation of variants product architectures. Sosa et al. (2003) observe the
that possess the same function with varying capacities importance of integrality and modularity in design team
(Simpson, Maier, & Mistree, 2001a). A generational interactions and introduce a method of identifying
platform leverages product life cycles for rapid next whether a system is modular or integral based on anal-
generation development (Martin & Ishii, 2002). One ysis of component interactions using a design structure
endeavor towards product platform development is to matrix (DSM). Fixson (2002) constructs a DSM to ana-
design product families in the way of “stretching” or lyze the total number of functions that components
“scaling” (Rothwell & Gardiner, 1990). under consideration provide on the other, based on
which modular and integral architectures are identified.
Oosterman (2001) investigates how to match product
Product architecture architectures with an organization in order to improve
product development projects. Whitney (2003) studies
The concept of architecture, with respect to product total modularity and interfaces in the context of design
design, is synonymous with the layout, configuration, or economy. Cutherell (1996) finds that integral architec-
topology of functions and their embodiment (Van Wie, tures are often driven by product performance or cost,
Rajan, Campbell, Stone, & Wood, 2003). Product archi- while modular architectures are driven by variety, prod-
tecture can be defined as the way in which the functional uct change, engineering standards, and service require-
elements of a product are arranged into physical units ments.
and the way in which these units interact (Ulrich and Jiao and Tseng (1999) assert that a product family
Eppinger, 1995). Fujita and Yoshida (2004) point out architecture involves systematic planning of modularity
one important characteristic to discern the architecture and commonality in terms of building blocks and their
of a family of products from that of a single product— configuration structures across the functional, technical
the simultaneous handling of multiple products. Erens and structural views. Zamirowski and Otto (1999) point
& Verhulst (1997) consider the functional and physi- out the necessity to develop the product architecture and
cal architectures for product families and describe them platform by synchronizing multiple views such as those
using a package of single product models. Yu, Gonz- from customer needs, functional structures and physical
alez-Zugasti, and Otto (1999) approach product archi- architectures. The leveraging of modularity and com-
tectures from a functional perspective by defining the monality in product family architecture development is
architecture based on customer demands. Ulrich (1995) also supported by Siddique, Rosen, and Wang (1998).
discusses the relationship between product architectures Muffatto and Roveda (2002) study multiple aspects of
and managerial problems related to product strategies. product architectures including functions, requirements,
The typology of product architectures suggests that technological solutions, product concepts, product strat-
the architecture can be either integral or modular egies and platforms, as well as production and assem-
(Muffatto & Roveda, 2002). Modularity has been well blies. To address the question of how differences in the
studied from many perspectives (Bi & Zhang, 2001; product architecture affect resource consumption dur-
Fixson, 2002; Gershenson, Prasad, & Zhang, 2003). ing the design phase, Eppinger, Whitney, Smith, and
Ulrich and Tung (1991) define five categories of mod- Gebala (1994) link the task structure of the design pro-
ularity, i.e., component swapping, component sharing, cess to the product architecture.
fabricate-to-fit, bus and sectional modularity. Pine (1993)
adds a sixth: mix modularity, which is frequently encoun- Product variety
tered in the painting and chemical industries. While most
extensions (Du, Jiao, & Tseng, 2001; Kusiak & Huang, Product variety is defined as the diversity of products
1996) are built upon these basic modularity types, the that a production system provides to the marketplace
current practice mostly refers to the product architec- (Ulrich, 1995). Du et al. (2001) discern functional vari-
ture as physical structures in terms of physical parts or ety from technical variety. While functional variety is
components (Henderson & Clark, 1990). most related to customer satisfaction, technical variety
J Intell Manuf (2007) 18:5–29 9

is relevant to manufacturability and costs. The two types It is commonality that entails the difference of the
of variety motivate two different design strategies. The architecture of product families from that of a single
design for “functional” variety strategy aims at increas- product. While modularity resembles decomposition of
ing functional variety and manifests itself through vast product structures and is applicable to describing mod-
research in the business community, such as product line ule (product) types, commonality characterizes the
structuring (Sanderson & Uzumeri, 1995), equilibrium grouping of similar module (product) variants of a spe-
pricing (Choi, Desarbo, & Harker, 1990), product posi- cific module (product) type that is characterized by
tioning (Kaul & Rao, 1995), and so on. To the contrary, modularity. Therefore, clustering is the main concern of
design for “technical” variety aims to reduce technical commonality. Corresponding to the three types of mod-
variety so as to gain cost advantages. Under this category, ularity, there are three types of commonality, namely
research includes variety reduction program (Suzue & functional commonality, technical commonality, and
Kohdate, 1990), design for variety (Martin & Ishii, 1997), physical commonality.
design for postponement (Feitzinger and Lee, 1997), The correlation of modularity and commonality is
function sharing (Ulrich and Eppinger, 1995), design for embodied in class–member relationships. A product
modularity Erixon and Ostgren (1993), reconfiguration structure is defined in terms of its modularity, through
for variety (Brabazon & MacCarthy, 2004), etc. which module types are specified. Product variants
derived from this product structure share the same mod-
Modularity vs. commonality ule types and entail different instances of every module
type. In other words, a class of products (i.e., prod-
The concepts of modules and modularity are central uct family) is described by modularity, whereas prod-
in constructing product architectures (Ulrich, 1995). A uct variants differentiate according to the commonality
module refers to a physical or conceptual grouping of among module instances.
components that share some characteristics. Modularity
is to separate a system into independent parts or mod-
ules that can be treated as logical units (Newcomb, Bras, Product portfolio and product family positioning
& Rosen, 1996). What is important in characterizing
modularity is the interaction among modules. Modules The original mindset of developing product families is
are identified in such a way that between-module (inter- to make wide variety of products available and letting
module) interactions are minimal whereas within-mod- customers “vote” on the shelf. This practice, however,
ule (infra-module) interactions may be high (Ulrich, not only seems to be wasteful and expensive, but also
1995); therefore, decomposition is the main concern of tends to constrain customers’ ultimate satisfaction, lead-
modularity. ing to mass confusion (Huffman & Kahn, 1998). Many
In addition, product family design achieves modu- marketing studies have been reported that some of the
larity from multiple viewpoints, involving functionality, product variants may be more preferred as expected,
solution technologies, and physical structures. Corre- while others, although they may be equally sound in the
spondingly, there are three types of modularity associ- technical sense, may not be favored by the customers.
ated with product families: (1) functional modularity, (2) The mismatch between expectation and achievement is
technical modularity, and (3) physical modularity. These mainly due to the diversity of customer requirements
three types of modularity are characterized by specific and preferences. Besides, not all existing market seg-
measures of module interaction from a particular point ments create the same opportunity for different com-
of view. As for functional modularity, the interaction is panies in the same industry due to the discrepancy of
exhibited by the relevance of FRs across different cus- their targets, strategies, technologies, cultures, etc. Pine
tomer groups. Each customer group is characterized by (1993) reports some common problems of companies
a particular set of FRs. A customer grouping lies only giving customers more choices than they actually want
in the functional view and is independent of the other or need. This essentially means a product family posi-
two views, meaning it is solution-neutral. In the techni- tioning problem—how to offer right product variety to
cal view, modularity is determined according to techno- the right target market.
logical feasibility of design solutions. The interaction is Product family positioning (i.e., product portfolio
thus judged by the coupling of DPs to satisfy given FRs planning) has been traditionally dealt with in the man-
regardless of their physical realization in manufactur- agement and marketing fields with emphasis on portfo-
ing. In the structural view, physical interactions derived lio optimization based on customer preferences where
from manufacturability become the major concern of the objective is to maximize profit, share of choices,
physical modularity. or sales. Consequently, measuring customer preferences
10 J Intell Manuf (2007) 18:5–29

among multi-attribute alternatives has been a primary the primary concern of optimal product design (Krish-
concern in marketing research. Among many methods nan & Gupta, 2001) or decision-based design (Hazel-
developed, conjoint analysis is one of the most popular rigg, 1998). In typical preference-based product design,
preference-based techniques for identifying and evalu- conjoint analysis (Green & Krieger, 1985) has proven
ating new product concepts (Green & Krieger, 1985). to be an effective means to estimate individual level
While many methods excel in determining optimal or part-worth utilities associated with individual product
near-optimal product designs from conjoint data, tradi- attributes. The conjoint-based searching for optimal
tional conjoint analysis is limited to considering input product designs always results in combinatorial opti-
from the customers only, rather than to analyze dis- mization problems because typically discrete attributes
tinct conjoint data from both customers and engineering are used in conjoint analysis (Kaul & Rao, 1995).
concerns (Tarasewich & Nair, 2001). Multi-attribute utility analysis is widely used to pre-
In the engineering community, product portfolio deci- dict composite utilities for any feasible product profile
sions have been extensively studied with particular focus constructed from the underlying attribute level part-
on the costs and flexibility issues associated with product worth utilities. It assumes that the utilities of multi-
variety and mix (Lancaster, 1990). Meyer and Lehnerd ple attributes are mutually independent (Wassenaar &
(1997) introduce a market segmentation grid to leverage Chen, 2001). This may not hold true for a product port-
product families across multiple market segments. Far- folio where the customer-perceived utility of a partic-
rell and Simpson (2001) apply the market segmentation ular attribute may change due to the availability of
grid to identify potential platform leveraging strategies other offerings (for example, comparing with counter-
for a line of flow control valves. Meyer and Utterback part attributes or levels). In addition, combining differ-
(1993) attempt to map the evolution of a product family ent individual attribute utility functions into a single
based on extensions and upgrades to a product platform. multi-attribute utility function inevitably involves multi-
Nevertheless, the effect of product lines on the profit attribute weighting and normalization. The weights are
side of the equation has been seldom considered (Yano determined based upon the rank ordering of alterna-
& Dobson, 1998). Few industries have developed an tives; however, a selected alternative may result from
effective set of analyses to manage the profit from vari- the underlying weighting method rather than the qual-
ety and the cost of complexity simultaneously in product ity of the alternative itself (Wassenaar & Chen, 2001).
portfolio decision-making (Otto, Tang, & Seering, 2003).
Markus and Váncza (1998) and Jiao & Zhang (2005) Product positioning
observe the importance of taking into account the inter-
action between the customer and engineering concerns Product positioning involves decisions about abstract
for product portfolio planning. In particular, portfolio perceptual attributes and customer heterogeneity (Kaul
decisions with customer-engineering interaction need to & Rao, 1995). The basic principle lies in the multidimen-
address the tradeoffs between the economy of scope in sional scaling of customer perceptions via factor anal-
profit from the customers and markets and the disecon- ysis, discriminant analysis or similarity scaling (Green
omy of scope in design, production, and distribution at & Krieger, 1985). A number of multidimensional scal-
the backend of product fulfillment (Yano & Dobson, ing-based algorithms have been developed, dependent
1998). Moreover, achieving a synergy of engineering con- upon the number of ideal points (individuals or seg-
cerns among products in portfolio planning is deemed ments) in the joint space (Kaul & Rao, 1995). Genetic
to be increasingly beneficial given those efforts in many algorithms have been proven to outperform most exist-
industries to improve coordination of design and manu- ing optimal positioning algorithms in dealing with the
facturing activities across product families and platforms choice set heterogeneity between the customer’s deci-
(Morgan, Daniels, & Kouvelis, 2001). The literature re- sion setting and variations in the size of the individuals’
lated to product family positioning generally stems from choice sets (Balakrishnan & Jacob, 1996).
three broad fields that are closely correlated: optimal On the other hand, many algorithms have been for-
product design, product positioning, and product line mulated with the attempt to improve the realism of
design. Each is discussed in the sections that follow. the customer choice setting. Discrete choice analysis is
widely used to identify patterns in choices that custom-
Optimal product design ers make among competing products (Ho & Tang, 1998).
It allows for the examination of the interaction between
While traditional design emphasizes more on the design- market shares and product features, price, service, and
ers’ perspective Tarasewich & Nair (2001), measuring promotion with respect to different classes of custom-
customer preferences in terms of expected utilities is ers. In general, a probabilistic choice model tends to
J Intell Manuf (2007) 18:5–29 11

provide better solutions and larger share projections for Scalable product family design
new product positioning.
The scalable approach was first proposed by Simpson et
al. (2001a). They introduce a product platform concept
Product line design
exploration method based on robust design principles
by minimizing the sensitivity of performance variations
Most literature on product line design tackles the opti-
in scaling factors. Messac et al. (2002a) integrate physical
mal selection of products by maximizing the surplus—
programming into the product platform concept explo-
the margin between the customer-perceived utility and
ration method to increase its effectiveness. Fellini et al.
the price of the product (Kaul & Rao, 1995). Other
(2002a) consider commonality as a constraint in prod-
objectives widely used in selecting products among a
uct family design. Nayak, Chen, and Simpson (2002)
large set of potential products within a target mar-
incorporate the platform selection problem into the com-
ket include maximization of profit, net present value,
monality and performance trade-off procedure by max-
a seller’s welfare, market share, and share of choices,
imizing commonality at the product family level while
to name but a few.
satisfying the design requirements of each individual
While numerous papers in marketing literature deal
product. Messac, Martinez, and Simpson (2002b) intro-
with the selection problem using various objectives orig-
duce a product family penalty function to the selection
inated from the profit, few of them explicitly model the
of right combination of common and scaling variables,
costs of manufacturing and engineering design (Yano &
where those design variables causing more performance
Dobson, 1998). Kim and Chhajed (2001) observe that,
loss are identified as scaling factors. Fellini et al. (2002b)
in addition to choosing which products to produce, one
employ sensitivity analysis to derive the penalty on per-
must also choose the process by which these products are
formance loss due to commonality. In the context of
manufactured. Park and Simpson (2005) articulate the
designing a family of aircraft, Willcox and Wakayama
benefits of different resource sharing methods related to
(2003) apply multidisciplinary optimization to gain
product platform design. Ramdas and Sawhney (2001)
insight into the effect of design variable scaling across
consider situations where the fixed cost of a component
multiple aircraft that carry different missions yet share
is shared by two products. Dobson and Yano (1994)
common parts.
allow for complex interactions by admitting per-prod-
Scalable product family design involves two basic
uct fixed costs, resources that can be shared by multiple
tasks (Simpson, 2004). The first one is platform selec-
products, as well as technology choices for each. Morgan
tion—to determine which design parameters that take
et al. (2001) examine the benefits of integrating market-
common values. While many existing methods assume
ing implications of product mix with more detailed man-
that the platform architecture is known a priori (Fujita
ufacturing cost implications, which sheds light on the
et al., 1999), some approaches determine platform vari-
impact of alternative manufacturing environment char-
ables along with scalable variables during optimization
acteristics on the composition of the optimal product
(Akundi, Simpson, & Reed, 2005; Dai & Scott, 2004).
line.
The subsequent task is to determine the optimal values
of common and distinctive variables by satisfying per-
formance and economic requirements. Most approaches
Platform-based product family design consider only a single product platform, where each plat-
form variable is shared across the entire product fam-
Corresponding to the scalable and modular product ily. This strategy excels in computational simplicity; but
platforms, there are two types of approaches to plat- may lead to a situation that some low-end products may
form-based product family design. One common app- be over-designed and certain high-end products may be
roach is called scalable (namely parametric) product under-designed (Dai & Scott, 2004). The other strat-
family design, whereby scaling variables are used to egy is to consider multiple product platforms in product
“stretch” or “shrink” the product platform in one or family design, such that design variables can be shared
more dimensions to satisfy a variety of customer needs by any subset of product variants within the product
(Simpson et al., 2001a). The other approach is referred to family (De Weck, Suh, & Chang, 2003) Multiple-plat-
as configurational product family design, which aims to form design enhances exploration of the solution space,
develop a modular product platform, from which prod- whereas sacrificing the computational efficiency (Seep-
uct family members are derived by adding, substituting, ersad, Mistree, & Allen, 2002).
and/or removing one or more functional modules (Du Simpson (2004) reviews optimization algorithms
et al., 2001; Ulrich, 1995). widely used for product family design. Linear and
12 J Intell Manuf (2007) 18:5–29

non-linear programming algorithms, such as SLP, SQP, method by applying the QFD matrix to modular analy-
NLP and GRG, are employed in many studies, in addi- sis and coin it as modular function deployment (MFD)
tion to derivative-free methods including genetic algo- with focus on the evaluation of module integration.
rithms, simulated annealing, pattern search, and branch Yu et al. (2003) apply the DSM as a tool to iden-
and bound techniques. Some specific algorithms and tify highly interactive groups of product elements and
heuristics are also available for use depending upon par- to cluster them into modules. Hölttä and Salonen (2003)
ticular solution frameworks, including decision-based compare three modularization methods using commer-
design (Li & Azarm, 2002), target cascading (Kokko- cial products. They reveal that the MFD method is the
laras, Fellini, Kim, Michelena, & Papalambros, 2002), least repeatable, whereas the computerized DSM me-
0–1 integer programming (Fujita et al., 1999), physical thod is the most repeatable, and the heuristic approach
programming (Messac et al., 2002a), and the compro- falls in between. Malmström and Malmqvist (1998) inte-
mise decision support problem (Simpson et al., 2001a). grate the DSM and MFD methods to tackle both techni-
When the design space is small, exhaustive search tech- cal and economical aspects in the early stages of product
niques (Hernandez et al., 2001) or orthogonal arrays architecture development. Stone et al. (2000) formulate
(Blackenfelt, 2000a) are always adopted to enumerate a set of heuristics for grouping functions to form a mod-
various combinations of parameter settings. In practice, ule. Hölttä et al. (2003) develop a five-step algorithm of
most problems involve a large number of options, which grouping and creating a dendrogram for finding com-
constitutes a typical combinatorial optimization prob- mon modules across products for platforming a product
lem. As a result, genetic algorithms are most advocated family. Salhieh and Kamrani (1999) employ a cluster-
to tackle the combinatorial nature of product family ing technique for identifying design modules. Otto et al.
design problems (Simpson, 2004). (2000) propose a framework for architecting a family
of products that share interchangeable modules. They
Configurational product family design define a modularity matrix for one family of products
from a manufacturer, allowing commonalties to be easily
The configurational approach to product family design identified. Gershenson et al. (2003) provide an extensive
is also frequently called module-based product family comparison of several DSM-based methods for identi-
design (Simpson, 2004). It is based on the develop- fying modular architectures.
ment of modular product architectures. As defined by (2) Interface standardization. The importance of stan-
Ulrich and Tung (1991), a modular product architecture dardized interfaces in product architectures is recog-
involves one-to-one mappings from functional elements nized by Meyer and Lehnerd (1997). Sanderson and
in the function structure to the physical components of Uzumeri (1995) suggest that the product family evolu-
a product, where decoupled interfaces between com- tion may have been restricted if clear and robust phys-
ponents can be specified. Ulrich (1995) points out that ical interfaces are not developed and defined carefully.
the modular product architecture allows each functional Sanchez (1994) discusses the need for distinct and stan-
element of the product to be changed independently dardized interfaces between the vital parts of the prod-
by changing only the corresponding component. This uct family. Sundgren (1999) explores the concept of
is advantageous to produce custom-built products from interface management in new product platform devel-
standard models. It also makes standardization possible, opment. Blackenfelt (2000b) studies robust design of
which is essential to achieve the economy of scale; there- interfaces in order to increase interface commonality
fore, using modular product architectures, variety can be among a variety of products. Ulrich and Eppinger (1995)
created by configuring existing building blocks. Salient point out that product architecturing is tantamount to
issues regarding configurational product family design interface definition.
include module identification, interface standardization, Martin and Ishii (2002) introduce a coupling index to
and architecture embodiment as discussed next. assess the amount of coupling between modular inter-
(1) Module identification/modularization. Erlandsson, faces to facilitate product family planning for multiple
Erixon, and Ostgren (1992) develop a method with three generations. Sosa et al. (2003) identify the distinction
major steps that help to identify product modules. In among interfaces from the energy, material, spatial and
their method, the right product specification is attained informational aspects. Hillestrom (1994) proposes a
by adopting quality function deployment (QFD). method that helps the designer clarify how interfaces
Module creation, interface analysis and module config- between modules influence module functions and how
uration are carried out through creating different mod- to select the best interface location. Mikkola and
ular structures according to the QFD matrix (i.e., the Gassmann (2003) introduce a modularization function
house of quality). Erixon and Ostgren (1993) extend this for analyzing the degree of modularity in a given product
J Intell Manuf (2007) 18:5–29 13

architecture by taking into account the number of com- introduced in order to achieve an optimal balance. Ger-
ponents, number of interfaces and substitutability factor shenson et al. (2003) develop a measure of relative
of a given product architecture. van Wie et al. (2001) modularity for modular product design. Mikkola and
address the embodiment issues of module interfaces in Gassmann (2003) assume that the degree of modular-
terms of the connections between modules. ity in a given product architecture is constrained by
(3) Architecture embodiment. Newcomb et al. (1996) the composition of its components. Allen and Carlson-
investigate the methods of reasoning about sets of prod- Skalak (1998) introduce some measures of modular-
uct architectures and how to perform configuration de- ity for conceptual design. Ulrich (1995) simply defines
sign of modular products. Kamrani and Gonzalez (2003) the function-to-component ratio for each product as
discuss design for modularity using a genetic algorithm- one modularity metric. Hölttä and Salonen (2003) pro-
based method. Tatikonda (1999) studies the planning pose a measure of modularity based on singular value
and execution of different types of projects within a decomposition of the binary DSM. Guo and Gershen-
product family platform series. Gonzalez-Zugasti et al. son (2004) develop a metric to measure product mod-
(2000) formulate the design of a platform-based prod- ularity using a component-to-component connectivity
uct family as a general optimization problem consider- matrix. Siddique and Rosen (2001) account for both
ing the constraints of individual product variants and functional and form issues in their partitioning method
constraints of the family as a whole. that involves combinatorics. Stone et al. (2000) develop
Agard and Kusiak (2004) apply data mining tech- product family and customer needs ratings for modules.
niques to product family design with emphasis on map- (2) Commonality. Kota et al. (2000) develop a mea-
ping specific functional requirements to a technical sure that captures the level of commonality in a prod-
structure. Allahverdi et al. (2002) optimize modular pro- uct family. With application to automotive underbodies,
ducts as a generalized subset selection problem subject Siddique, Rosen, and Wang (1998) propose to measure
to a balance of the quality loss due to modularization component commonality and connection commonality
and the cost of reconfiguration. Chandrasekaran, Stone, in order to capture characteristics of platform common-
and Mcadams (2004) propose a template approach to ality and product variety. Maupin and Stauffer (2000)
product platform-based design based on prior product take into account simplicity, direct costs, and delayed
design knowledge and the flow and conversion of energy differentiation for commonality metrics. Thonemann and
through a product. Kurtadikar and Stone (2003) use high Brandeau (2000) present an approach for determining
level customer needs alone to define the product’s base the optimal level of commonality in a sub-product that
platform and differentiating modules during the con- does not differentiate models from the customer’s point
ceptual design stage of product development and thus of view. Emphasizing on component sharing, Ramdas et
plan a product family before considering any architec- al. (2003) present a methodology for determining which
ture. Robertson and Ulrich (1998) articulate three tools version of a set of related components should be offered
for supporting platform planning: product plan, differ- to optimally support a defined finished product port-
entiation plan and commonality plan. folio. In the work of Fellini et al. (2002a), an optimal
design problem is formulated as the maximization of
Metrics for product family design commonality by choosing the product components to be
shared without exceeding a user-specified performance
Product family design essentially entails a type of multi- loss tolerance and subject to different levels of perfor-
objective optimization problems (Simpson, Siddique, & mance losses. McAdams and Wood (2002) develop a
Jiao, 2005). In many cases, such multiple criteria deci- quantitative metric for design-by-analogy based on the
sion-making, given a number of alternatives at differ- functional similarity of products. Thevenot and Simpson
ent levels of abstraction of the product architecture, (2005) compare various commonality indices for assess-
requires to leverage on three pillars: cost, revenue and ing product families, including the Degree of Common-
performance. In addition, it needs to weigh the reve- ality Index (Collier, 1981), Total Constant Commonality
nue from product cannibalization by commonality with Index (Wacker & Trelevan, 1986), Product Line Com-
respect to the cost savings from commonality (Robert- monality Index (Kota et al., 2000), Percent Common-
son & Ulrich, 1998). ality Index (Siddique et al., 1998), Commonality Index
(1) Modularity. Prasad (1998) studies the product and (Martin & Ishii, 1997), and Component Part Common-
process complexity associated with design for variety, ality Index (Jiao & Tseng, 2000).
highlighting the importance of determining the right (3) Variety/distinctiveness. Martin & Ishii (1997) quan-
amount of decomposition. To quantify such a granu- tify the costs of providing variety in order to quantita-
larity paradox, a measure of communication effort is tively guide designers in developing products that incur
14 J Intell Manuf (2007) 18:5–29

minimum variety costs. Through commonality analysis, Nelson, Parkinson, & Papalambros, 2001). Many stud-
van Wie et al. (2006) study how differences between ies have revealed that such a profit measure based on
platform elements and differentiating elements are evi- the dollar value is unrealistic in most cases (Tarasewich
denced in the product layout or configuration. Simpson & Nair, 2001). As such, researchers have been devel-
and D’Souza (2004) introduce a genetic algorithm-based oping various instruments to improve the measurement
approach to product family design that balances the of profit performance. Balakrishnan and Jacob (1996)
commonality of the products in the family with the indi- introduce share of choices as the objective. Chen and Lin
vidual performance (i.e., distinctiveness) of each prod- (2002) try to maximize performance ratings of design
uct in the family. A Product Variety Index is proposed concepts while minimizing functional coupling and cou-
by Simpson et al. (2001b) to help resolve this tradeoff. pling of design concepts with respect to constraints.
Dobrescu and Reich (2003) propose a variety index and Michalek et al. (2005) formulate the evaluation prob-
a standardization index that resemble the commonality lem as profit maximization by minimizing the technical
indices of Martin and Ishii (2002). performance deviation. De Weck et al. (2003) propose to
(4) Cost. Kim and Chhajed (2001) develop an eco- optimize product platform design by maximizing overall
nomic model that considers a market consisting of a high product family profitability and reducing the develop-
segment and a low segment. They determine that large ment time and cost. Jiao & Zhang (2005) combine the
commonality decreases production costs but makes the consumer surplus with the producer surplus based on
products more indistinguishable from one another, the customer-perceived value of design.
which makes the product more desirable for the low Typical approaches to estimate costs and values coin-
segment but less desirable for the high segment. Fisher cide with the traditional principle of capital budgeting
et al. (1999) present an analytic model of component that is based on discounted cash flows (DCF) analy-
sharing based on empirical testing of varying practice sis. When dealing with numerous options associated
of component sharing for automotive braking systems. with product family design, the DCF approach tends
Fujita and Yoshida (2004) develop a monotonic cost to underestimate the upside potentials to a design pro-
model for the assessment of benefits of commonality. ject from management flexibility (Kogut & Kulatilaka,
Gonzalez-Zugasti et al. (2000) propose a methodology 1994). Real options have been applied to value specific
to design product platforms and variants with consider- aspects of product development, such as design modu-
ation of technical performance requirements and prod- larity (Baldwin & Clark, 2000). Otto et al. (2003) explore
uct family costs. the real options concept for determining proper levels of
Fixson (2005) outlines a roadmap for product archi- independent product architectural attributes. Jiao et al.
tecture costing from a product life cycle perspective. (2005a) apply a real options framework to flexibility val-
Park and Simpson (2005) examine the effects of com- uation of a product family architecture. Banerjee and de
monality decisions on individual costs based on activity- Weck (2004) propose a flexible product option valuation
based costing. To link modularity and the cost, Fixson framework to evaluate the conditions where a flexible
(2002) develops a multi-dimensional product architec- architecture is no longer financially viable vis-à-vis fixed
ture description method that considers the level of func- architectures. The rationale of combining technologi-
tion-component allocation, interface intensity, interface cal and business decisions in real options valuation for
reversibility, and interface standardization. Siddique and optimal design decisions is supported by Georgiopoulos
Rosen (2001) assess the cost implications of product et al. (2002). Jiao et al. (2005a) propose a hybrid real
architectural decisions when product architectures allow options model considering the value of flexibility either
sharing of parts, modules, or components of a product inherent in a product family design project or that can
across product families. The savings from the reuse of be built in product platforms.
designs affect both development cost and time (Sid- (6) Platform-related Metrics. Meyer and Lehnerd
dique, 2001). To design robust product families and (1997) develop two platform related measures, named
determine product platform extent, Hernandez et al. platform efficiency and platform effectiveness, for eval-
(2001) estimate production time, the material cost, and uating the performance of product families. Focusing
the inventory cost by modeling manufacturing issues on the generational aspect of product platforms, Mar-
as functions of design variables. Martin & Ishii (1997) tin and Ishii (2002) develop two indices, called Gener-
employ a setup cost index as a surrogate for the indirect ational Variety Index and Coupling Index, to measure
cost associated with product variety. a product’s architecture. Dahamus et al. (2001) propose
(5) Profit/valuation. Numerous methods dealing with to architect a portfolio of products by exploiting com-
optimal design use various objectives originated from monality and reusing modules across the family of prod-
the profit or expected revenue (Fujita & Yoshida, 2004; ucts, rather than a fixed product platform upon which
J Intell Manuf (2007) 18:5–29 15

derivative products are created. That is, the platform nize and manage product knowledge through a knowl-
itself allows for several possible sizes and types. Fellini edge component that includes configuration rules and
et al. (2002b) introduce a metric, called Sharing Penalty constraints. Bohm and Stone (2004) investigate the rep-
Vector, to the optimal selection of product platforms resentation of functionality for supporting reuse. Shar-
for family products with mild variation. Messac et al. man and Yassine (2004) study some forms of abstraction
(2002b) introduce a Product Family Penalty Function to for describing product architectures, including DSM,
promote platform commonality during product family molecular diagrams, and visibility-dependency signature
optimization. De Weck et al. (2003) adopt a market seg- diagrams. Costa and Young (2001) introduce a prod-
ment model using the sales volume, the price and the uct range model (i.e., product families) for information
competing product alternatives for product family and modeling of variant and adaptive design. Tiihonen et
platform portfolio optimization. Jiao & Tseng (2004) al. (1998) develop a method of managing and model-
develop a design customizability index and a process ing a product family as a configurable product, which
customizability index for evaluating the cost effective- is based on the conceptualization of components, attri-
ness of a design to be customized in order to meet indi- butes, resources, ports, contexts, functions and
vidual customer needs. Zha et al. (2004) introduce two constraints. Jiao et al. (1998) observe different data
metrics, market efficiency and investment efficiency, to types underlying product families that involve product-
the evaluation and selection of product design for mass to-product, product-to-family and family-to-family rela-
customization. tionships. To characterize variety and its derivation, Jiao
et al. (2000) propose a generic variety structure consist-
Product family modeling ing of common product structures, variety parameters,
and configuration constraints.
Baldwin & Clark (2000) develop a discipline-indepen- Shooter et al. (2005) develop an information man-
dent data model to provide constructs for modeling agement infrastructure for product family planning and
products with optional contents. Felfernig, Friedrich, platform customization. Corbett & Rosen (2004) develop
and Jannach (2001) apply the unified modeling language computational methods for modeling and combining
to the modeling of configuration knowledge bases for multiple design spaces so that product family configu-
mass customizable products. The initiative of Product rations can be identified while satisfying all constraints.
Family Classification Tree emphasizes the classification To articulate design of modular product architectures,
of end-products and/or modules from a functional view- Rosen (1996) considers three design spaces in terms of
point (Bei & MacCallum, 1995). To facilitate repre- material compatibility, connections and covers. Schwarze
sentations from multiple perspectives, Generic Product (1996) studies the fundamental problems of configur-
Modeling is advocated to represent product families ing multiple-variant products with emphasis on con-
from both commercial and assembly views (Wortmann figuration information modeling. Johannesson (1992)
et al., 1997). McKay et al. (1996) describe product fam- attempts to build computer-based parametric product
ilies from both sales/customer and assembly views by family models. Computer supported product platform
incorporating descriptions of detailed product data into models are proposed by Claesson, Johannesson, and
specifications of product variety. Siddique and Rosen Gedell (2001). The work of Sivard (2000) emanates from
(1999) develop a graph grammar approach to product axiomatic design and the ISO standard STEP AP214
platform design. Du et al. 2002a (2002a) apply graph with the purpose of improving computer representation
grammars to product family modeling. A graph rewrit- of information that concerns product families and prod-
ing system is developed to support product family con- uct platforms. Johannesson and Claesson (2005) report
figuration design (Du, Jiao, & Tseng, 2002b). Set-based a combined function-means and parametric modeling
model is an attempt to the formal representation of approach to product platform design. Their concepts are
product platform design and manufacturing processes being deployed by Saab Automobile as the next logical
(Finch, 1999). Männistö (2000) studies the conceptual evolution of product description systems.
modeling of product families, with particular empha-
sis on the problems related to the evolution of product Product family design support systems
family descriptions and the product individuals created
according to them. Kusiak and Huang (1996) and O’Grady and Liang (1998)
van Wie et al. (2003) consider product architecture put forth design with modules that centers around mod-
representation as organizing a deluge of information in ule selection. Liang and O’Grady (2000) focus on a
terms of both function and form. To model product fam- particular design environment where modules may be
ily configuration, Zhang et al. (2005) propose to orga- available from one or more geographically dispersed
16 J Intell Manuf (2007) 18:5–29

sources, and where data concerning the modules may be actually committed and product quality and lead times
in a multitude of databases scattered across the globe. are determined per se. For a given design, the actual cost
Huang and Liang (2001) develop formalism for design depends on how the production is planned and to what
with modules, such that customer requirements are met extent the economy of scale can be realized within the
using modules from suppliers geographically separated existing manufacturing capabilities. This implies that the
through diverse computer platforms. claimed rationale of product family design can only be
Siddique and Rosen (2001) develop a product family fulfilled at the production stage (Jiao & Tseng, 2004).
configuration reasoning system to identify common plat- The direct consequence of product customization on
forms from a collection of similar existing products, as production is observed as an exponentially increased
well as to generate product families from these common number of process variations (referred to as process
platforms. Siddique and Boddu (2004) propose an infor- variety), such as diverse machines, tools, fixtures, set-
mation framework of integrating customers into config- ups, cycle times, and labor (Wortmann et al., 1997).
uration design using a graph grammar and templates for Process variety is embedded as a set of generic items,
modules. Zha et al. (2004) examine the development of a each of which instantiates some similar items of the
knowledge-intensive support scheme and a comprehen- same type, in addition to various constraints regarding
sive fuzzy clustering and ranking methodology for prod- re-sequencing of operations processes (namely process
uct family design evaluation. Forza and Salvador (2002) structural changes). Process variety introduces signifi-
investigate how product configuration systems support cant constraints to production planning and control,
the order acquisition and fulfillment process in a high e.g., preventing make-to-order systems from achieving
product variety environment. Huang (2004) develops cost-effective customization capabilities (Kolisch, 2000).
an agent-based knowledge management system for deci- Regardless of the negative impact of process variety, the
sion support of modular product collaborative design. common components and the same basic product struc-
Online product configurators have recently received ture assumed by the customized products in a family
much attention to enable customers to interactively inherently enable similarity in the corresponding pro-
specify and adapt a product according to their individ- duction processes (Martinez, Favrel, & Fhodous, 2000).
ual preferences (Sabin & Weigel, 1998). Bramham and Similar to a product family, a process family comprises a
MacCarthy (2003) examine the empirical evidence of set of similar production processes that share a common
available configurators in terms of matching configura- process structure (referred to as a process platform).
tor attributes against business strategies. Hvam (2004) Consequently, companies are interested in configuring
reviews the design and implementation of product con- existing operations and manufacturing processes (termed
figuration systems from the viewpoint of industrial as production configuration) by exploiting similarity
applications. Simpson et al. (2003) investigate the frame- among product and process families so as to take advan-
works for web-based platform customization. Common tage of repetitions (Schierholt, 2001). Besides leveraging
configuration systems for product families necessitate the cost of delivering variety, exploiting process families
product specific knowledge and often overstrain cus- around process platforms can reduce development risks
tomers (Blecker et al., 2004a). Advisory systems are thus by reusing proven elements in a firm’s activities (Sawh-
advocated to guide customers according to their profile ney, 1998).
and requirements through a personalized configuration
process ending with the generation of product vari- Managing process variety
ants that better fulfill the real customer needs (Blecker,
Abdelkafi, Kreutler, & Friedrich, 2004b). Blecker et al. Ahmad and Schroeder (2002) observe that, when a firm’s
(2004a) present a multi-agent based design for the con- product structures or product mixes change, its process
figuration process of product variety. Ardissono et al. structure should also be changed to balance flexibility
(2003) report on an EU-funded project, CAWICOMS and efficiency. The traditional approach to deal with a
Workbench, which aims at next generation web-based large number of variants associated with product fami-
applications that support distributed configuration of lies is to treat each product as an individual bill-of-mate-
products and services within a supply chain. rials (BOM), which however leads to a data explosion
problem (Olsen & Stre, 1996). To overcome the limi-
tations of traditional BOMs in handling a large num-
Manufacturing and production for product families ber of variants, the generic BOM (GBOM) concept is
developed by van Veen (1992). The GBOM defines a
While seeking technical solutions is the major concern in generic product as a set of variants that can be iden-
design, it is at the production stage that product costs are tified through specifying alternative values for a set of
J Intell Manuf (2007) 18:5–29 17

parameters (Hegge & Wortmann, 1991). The generic involve organizational, strategic, technology and cost
bill-of-materials-and-operations is put forth by Jiao et related issues. Tsubon eet al. (1994) study the relation-
al. (2000) by unifying BOMs and routings to accom- ship between component part commonality and manu-
modate large numbers of product and process variants. facturing flexibility. Siddique, Rosen, and Wang (1998)
For multiproduct and multiprocess production systems, and Wilhelm (1997) demonstrate that the level in the
Aydiny and Gugor (2005) develop a relational database product hierarchy at which commonality is pursued var-
approach to generating BOMs and executing MRP. ies with respect to the deployment of production pro-
Focusing on reducing subassembly proliferation and cesses.
the cost of offering product variety, Gupta and Krishnan Kusiak (2002) examines the integration of modular-
(1998) propose a methodology for designing product ity in terms of products, processes, and resources. Sako
family-based assembly sequences. De Lit and Delcham- and Murray (2000) present a holistic view of modularity
bre (2003) put forward the integrated design of product in design, production and use within the context of the
families and the corresponding assembly systems. He global automotive industry. The impact of product mod-
and Kusiak (1996) discuss the design of assembly sys- ularity on the design for assembly and product life cycles
tems for modular products by dividing an assembly line is discussed in He and Kusiak (1996) and Newcomb et al.
into the basic and variant subassembly lines, such that (1996). Bi and Zhang (2001) present taxonomy on mod-
the basic subassembly line is used for common and basic ularity technologies with applications to manufacturing.
operations, whereas the variant ones for variant opera- Salvador et al. (2002) explore how manufacturing char-
tions. Fujimoto and Ahmed (2003) investigate assembly acteristics affect the appropriate type of modularity to
process design for handling variety with focus on how be embedded into a product family architecture.
to manage product and process-based manufacturing Meyer and Dalal (2001) study the management of
complexities of an entire product family. Klocke et al. platform architectures along with manufacturing pro-
(2000) develop systematic methods for product-related cesses. The concept of a process platform is introduced
modular process design at different production planning by Jiao et al. (2003) for product and process variety
and optimization phases. Lee and Tang (1998) develop management. To support computerized process deriva-
a model for evaluating the costs and benefits associated tion, the group technology is applied to construct coding
with a situation in which the product differentiation can schemes for representing process variations in relation
be delayed through product and process redesign. to product variants (Jiao, Zhang, & Pokharel, 2005c).
Romanowski and Nagi (2002) combine the data-mining
Process platform and graph-theoretic approaches in order to generate
generic BOMs incorporated with design and manufac-
Meyer and Lehnerd (1997) expand the common defini- turing information. To construct a process platform, Jiao
tion of a platform to include possible commonality in et al. (2005b) employ the text mining and graph match-
processes and production. In particular from the pro- ing techniques to identify generic routings from large
duction and assembly perspectives, a platform implies amount of production information and process data.
a focus on commonality of production tools, machines Further to coordinate product and process variety, Jiao
and assembly lines (Sanchez, 1994). As a consequence, et al. (2005b) adopt the association rule mining tech-
some companies in the automotive industry have con- nique to identify the interrelationships between product
sidered it more interesting to define a platform on a variants and process variations.
manufacturing-assembly basis rather than on a product
development basis, so as to better exploit commonality Production configuration
among production process (Wilhelm, 1997).
The benefit of designing product families comes from Stobaugh and Telesio (1983) represent early efforts to
a reduction of components in inventory and compo- develop configuration models for manufacturing. They
nent handling, reduction of component types and man- purport to boil the manufacturing task list down to
ufacturing processes, and increased production volumes a few “generic tasks” and then address major deci-
(Fisher et al., 1999). However, sharing components in sions through “choices within each type”. Hill (1994)
a product family may lead to a lack of distinctiveness, describes five generic process types across 25 dimen-
and shared components in one product often exceed the sions with an intention to capture product and market
requirements of other products, which causes additional requirements, manufacturing characteristics, investment
production costs (Krishnan & Gupta, 2001). Nobelius and cost issues, and infrastructure choices. Hayes and
and Sundgren (2002) point out that the potential mana- Wheelwright (1984) develop a product–process matrix
gerial difficulties associated with the part sharing process to address internal fit at the process choice level.
18 J Intell Manuf (2007) 18:5–29

Bozarth and McDermott (1998) examine the configura- process and supply chain design decisions. Fine et al.
tion research in the broad area of manufacturing strat- (2005) purport to expand traditional concurrent engi-
egies and conclude that configuration models are well neering to three dimensions, namely 3D-CE, in terms
suited to studying complex multivariate organizational of products, processes, and the supply chain. He argues
phenomena. that all three domains possess certain architectures, and
Schierholt (2001) presents the concept of process con- thus simultaneously considering these architectures is
figuration that combines the principles of product con- the key to accommodate low volume, large variety, and
figuration and process planning. In accordance with a high-speed operations. Jiao, Huang, and Tseng (2004)
given product family, a process family, consisting of a set discuss concurrent enterprising for mass customization,
of process variants, is concerned with the fulfillment of which aims to align customers, products, processes, and
all product variants in the family. Commonality across logistics for delivering an increasing product variety with
the variety of product variants leads to a number of same reasonable costs. McKay and Pennington (2001) provide
or similar operations, processes and sequences among an operational level framework for modeling products,
process variants (Schierholt, 2001). In reality a domi- processes and supply chains. Fixson (2005) utilizes the
nant flow direction usually presents in a general flow product architecture as a coordination mechanism for
shop (Oosterman, Land, & Gaalman, 2000). Therefore, decision-making across the product, manufacturing pro-
a common product structure and a common process cess, and supply chain domains. Garg (1999) develops a
structure exist within a product family, and variety is supply chain modeling and analysis tool for designing
embodied in different variants (instances) of these com- products and processes within a supply chain frame-
mon structures. work. Singhal and Singhal (2002) present a compatibil-
To support production configuration, Jiao, Zhang, ity matrix for identifying desirable product ideas with
and Prasanna (2005d) study the modeling of process consideration of operations and marketing capabilities,
variety using object-oriented Petri-Nets with change- along with the supply chain, the product and the pro-
able structures. Siddique (2005) develops a method to cesses used to manufacture the product.
explicitly consider existing assembly plant configuration
and resources during selection of assembly processes for Sourcing/outsourcing
new product family members. Molina et al. (2005) argue
that the next generation manufacturing systems must Salvador et al. (2002) investigate multiple companies
be able to provide increased levels of flexibility, recon- and their product groups and find that the chosen type
figurability and intelligence to allow them to respond to and degree of product modularity have implications on
product variety. Mehrabi, Ulsoy, and Koren (2000) envi- where to source components. Appelqvist and Heikkilä
sion reconfigurable manufacturing systems as a new par- (2003) propose a model for product, process, and sup-
adigm in manufacturing, in which it is possible for rapid ply chain configuration in a build-to-order environment.
adjustment of production capacity and functionality in They argue that product standardization makes it pos-
response to increased variety of products. Benjaafar, sible to move the order penetration point downstream,
Heragu, and Irani (2002) discuss the flexible, modular whilst downstream order penetration makes it possi-
and reconfigure layout configurations for meeting the ble to effectively differentiate by providing fast, flexible
needs of multiproduct enterprises. deliveries to customers. Momme, Moeller, and Hvolby
(2000) observe that the level of modularity affects a
firm’s sourcing strategy—outsourcing commodity items
Supply chain issues of product families while keeping components of strategic relevance in-
house. Chakravarty and Balakrishnan (2001) analyze
Customer requirements for flexibility, agility, cost the minimum number of module options that are req-
efficiency and product variety force companies to recon- uired for a desired amount of variety.
figure their supply chains and to focus more on collab- Dahmus and Otto (2001) observe that the more the
oration with external partners (Salvador et al., 2002). product architecture enables late customization, i.e.,
The design of a supply chain has an essential influ- postponement, the more it can contribute to savings
ence on how a manufacturer of complex products orga- in inventory and work-in-process costs. This also sug-
nizes and coordinates the stream of innovative products gests that the use of common components allows
through platform and architectural design strategies vis- inventory levels to be lowered through risk pooling
à-vis the sourcing, manufacturing and distribution strat- (Gerchak, Magazine, and Gamble, 1988). From a sup-
egies (Mikkola & Skjøtt-Larsen, 2004). Rungtusanatham ply chain perspective, risk pooling is one of the most
& Forza (2005) advocate to coordinate the product, noticeable effects of product platforms and component
J Intell Manuf (2007) 18:5–29 19

commonality across product families. Gupta and Krish- crucial processes, through which end products embody
nan (1999) study component sharing with multiple com- specific functional features. This implies to decouple the
ponent types when considering the economy of scale in fulfillment process into standardized and customer spe-
procuring one unique component type instead of multi- cific parts. Design for postponement has been put forth
ple component types from a single supplier. Their find- in line with the product architecture and modularity
ings coincide with the wisdom of reducing complexity of principles (Lee & Billington, 1994).
a product family through product design by leveraging Lee and Tang (1998) point out that product modular-
common characteristics among the products within the ity affects the extent to which strategies for postpone-
family. ment and late customization can be realized, and thus
Collier (1981) identifies part commonality as a way maximizing the benefits from a postponement strategy
to reduce the safety stock level for a given service level. may require re-sequencing of processes. Pagh and Coo-
Baker, Magazine, and Nuttle (1986) show that the safety per (1998) distinguish between manufacturing, logis-
stock of unique components increases if a certain service tics and full postponement. Zinn and Bowersox (1988)
level is to be maintained, while the stock for a common label two types of postponement. Time postponement
part may be lower compared with the unique parts it refers to postponement of the delivery of the prod-
replaces. Gerchak, Magazine, and Gamble (1988) draw ucts until the customer orders arrive. Form postpone-
similar conclusions for an arbitrary number of prod- ment means delaying the differentiation of the products
ucts as long as the costs of the product-specific com- to later stages of the supply chain. While the former
ponents that are replaced by a common one are the excels in meeting the increased customized demands
same. Furthermore, product architectures can reduce with lower inventory levels, the latter improves the lead-
increased demand volatility from upstream, i.e., the bull- time performance. Su, Chang, and Ferguson (2005) eval-
whip-effect, if they allow for parallelization of produc- uate the time postponement and form postponement
tion to achieve short lead times (Perera, Nagarur, & strategies for implementing large product variety. They
Tabucanon, 1999). argue that the time postponement structure maps to
With focus on the automotive industry, Fredriksson a make-to-order supply chain, whereas the form post-
(2002) analyzes the influence of organizational forms on ponement supply chain structure fits into a hybrid make-
the outsourcing decision of modular assembly units in to-stock and make-to-order environment, which are
relation to product modularity. Takeishi and Fujimoto dedicated to common components and customized prod-
(2003) study the impact of modularization on the prod- uct offerings, respectively.
uct, production and supplier systems in the car industry.
A number of studies have emphasized on the synchroni- Resource allocation
zation of supply chain management and product design
decisions (Lee & Sasser, 1995). Such recognitions are In a networked manufacturing supply chain, allocat-
consistent with the trend towards outsourcing both man- ing proper production volumes to various manufactur-
ufacturing activities and design activities (McCarthy & ing sites is a complex task involving diverse markets,
Anagnostou, 2004), as well as the importance of sup- product families, manufacturing capabilities, as well as
plier involvement in product development (McIvor & final-production and sub-assembly supplies. Based on
Humphreys, 2004). Petersen, Handfield, and Ragatz product family principles, Ng and Jiao (2004) study the
(2005) argue that integrating suppliers into the prod- tradeoffs of resource utilization, inventory volumes and
uct development process has direct impact on manu- work-in-process levels. To support resource allocation
facturing process design decisions and on supply chain across product, process, and supply chain configura-
configuration decisions. tions, Blackhurst, Wu, and O’Grady (2005) develop a
network-based product chain decision model for describ-
Decoupling/postponement ing supply chain operations. Huang et al. (2005) recog-
nize the issue of sharing manufacturing resources and
Product family development implies a make-to-order or supply sources by exploiting common components and
assemble-to-order production, which means decoupling modular product structure to be shared across the indi-
and delaying certain stages of product fulfillment. The vidual products. Chong, Ho, and Tang (1998) point out
postponement strategy means that some of the activities that product proliferation and variety lead to increased
in the supply chain, may it be production, assembly, or supply chain complexity, unacceptably high production
even design, are not performed until customer orders are and inventory costs, as well as long time to market.
received (van Hoek, 2001). Feitzinger and Lee (1997) Thonemann and Bradley (2002) investigate the impact
stress that postponement is about delaying the timing of of product variety on supply chain performance. They
20 J Intell Manuf (2007) 18:5–29

determine that product variety has significant effect on much to be desired in order to achieve system-wide
supply chain lead time especially when setup times are solutions for product family design and platform-based
significant; therefore, it becomes imperative to adjust product development. In this regard, the following areas
the decision variables and parameters related to manu- are suggested as the avenue for further research efforts.
facturing processes and supply chains in order to improve
performance under high product variety. Customer integration for product families

Supply chain configuration The driving force behind product family design and
development is the enterprises’ positioning of custom-
According to demand chain management principles, the ers at the center of value creation and involving cus-
preferences of end customers should be the basis for tomers into the product fulfillment process. Of primary
configuring supply chains (Lee & Sasser, 1995). Park, importance in product families is the interaction with
Ghosh, and Murthy (2000) present a comprehensive customers (Blecker et al., 2004a). On the technical side,
model for integrated product platform and global sup- designers have always assumed that customers’ satis-
ply chain configuration. Huang et al. (2005) analyze the faction with the designed product is sufficiently high as
impact of platform products, with and without common- long as the product meets the prescribed technical speci-
ality, on decisions pertaining to supply chain configura- fications; however, what customers appreciate is not the
tion and the consequent performance of the configured enhancement of the solution capability but the func-
supply chain. Kim, Leung, Park, Zhang, and Lee (2002) tionality of the product. This means that the traditional
develop a mathematical model and a solution algorithm dimensions of customer satisfaction may deserve scru-
to assist the manufacturer in configuring its supply chain tiny, for example, identifying those product character-
for a mix of multiple products sharing some common raw istics that cause different degrees of satisfaction among
materials and/or component parts. customers; understanding the interrelation between the
Based on the concept of ontology-oriented constraint buying process and product satisfaction; determining the
networks, Novak and Eppinger (2001) find statistically optimal amount of customization and customer inte-
significant relations between supply chain structures and gration; explaining the key factors regarding the value
product architectures for luxury and high-performance perception of customers; and justifying an appropriate
vehicles. Choi and Hong (2002) classify supply chains number of choices from the customers’ perspective.
according to formalization, centralization and complex- Equally important are customers’ decision-making
ity, and thus suggest that integral supply chains tend to processes when interacting with product families and in
be characterized by high levels in all three dimensions. turn developing proper fulfillment capabilities. At the
Fine et al. (2005) conjecture that integral products would end providing decision support to customers is deemed
ideally be built by integral supply chains, whilst modu- to be important. This coincides with consumer behaviors
lar products tend to be produced by modular supply in business systems based on customer involvement in
chains. Based on examination of the automotive indus- the product customization process (Huffman & Kahn,
try, Doran (2003) infers that the general shift toward 1998). While most product family-based customization
more modular product designs in that industry forces approaches implemented in practice are based on offer-
suppliers to reevaluate their strategies. As a result, many ing a huge number of variety and choices, the perception
would-be first-tier suppliers must now position them- of choice and the joy or burden of configuration experi-
selves as value-added second-tier suppliers, in order to enced by customers are not well understood. Many ques-
fit with the more flexible characteristics of the emerging tions are pending. For example, what are the incentives
modular supply chains. for integrating customers into value creation? What fac-
tors drive customers to interact with a configurator?
How many variants should be explored and changed
Prospects for future research before making a final decision? Are there any specific
patterns that customers follow when interacting with a
As witnessed in this comprehensive review, product fam- product family design system? And how do different lev-
ily design and development has been tackled from a els of the decoupling point (i.e., where the customer is
broad scope of product fulfillment, including both front- integrated into value creation) influence customer inte-
end and back-end issues. Owing to the flurry of var- gration and how does this affect the performance of a
ious research efforts, this emerging field of research product customization system? Toward this end, prod-
has matured rapidly in the past decade (Simpson, 2004; uct family development needs to be incorporated with
Simpson et al., 2005). From a holistic view, there is still more marketing engineering decisions (Michalek et al.,
J Intell Manuf (2007) 18:5–29 21

2005), as well as customer perceptions (Blecker et al., the support of product families and e-commerce tech-
2004b). nologies?

Product family design support Manufacturing and logistics issues

Product family-based customization has often been The premise of product family design and platform-
enacted as an “after thought” from the manufacturer’s based product development relies on the belief that
point of view, and thereby only limited options are per- flexible manufacturing systems along with scalable and
ceived by the customers. The prevailing principle is a modular product structures can significantly reduce the
two-stage process of product development. While prod- fixed cost in comparison with the variable cost, allow-
uct architectures and the range of possible variety are ing variety to be provided without incurring significant
determined a priori during a product family architect- cost increases. Nonetheless, flexible manufacturing and
ing stage, a subsequent design and development stage the corresponding planning systems are necessary but
takes place in close interaction between the customers insufficient for successful implementation of product
and the manufacturer. Based on what has been learned families. These systems have to be supplemented by
from the second stage, the product family architecture information technologies capable of handling the infor-
and platform may be redesigned or upgraded, which mation flows and transaction costs involved in the ful-
in turn leads to capability enhancement on the manu- fillment of product families. The advent of pervasive
facturer side. One of the important research topics is connectivity by the Internet provides the necessary and
product architecture and platform modeling. Compar- affordable connections among all parties in a product
ing with numerous efforts in product family optimiza- family-based customization system. As a result, the suc-
tion design, this field however has so far received least cessful implementation of product families depends on
attention and little achievement has been reported (Cor- the extension of concurrent engineering beyond the
bett & Rosen, 2004). It is imperative to call for rigorous traditional boundary of design and manufacturing to
research that synthesizes useful ingredients from those include customer interaction, marketing, service and
establishments in the artificial intelligence field such as recovery. It has been suggested that concurrent enter-
configuration topology, software product families, and prising is very much in line with the idea of the real-
architectural modeling. time economy where the customers are central to the
Although the basic principles of product family design value creation (Prasad, 1998). More research efforts are
are understood and well documented in the literature, expected for building up rigorous frameworks of recon-
quite a few fundamental issues require further exami- figurable processes and process platforms, integrated
nation. Some pending issues include, for example, the information management for product and process fam-
difference of customer-perceived variety from techni- ilies, coordination of product and process variety, etc.
cal variety in fulfillment, the optimal degree of product Another important aspect is for the enterprise to cre-
differentiation, the mechanisms of interrelation between ate capabilities with dynamic stability that enable the
modularity and commonality, implications of adaptabil- firm to adapt, integrate, and reconfigure the manufac-
ity, flexibility, reusability and customizability, product turing skills and competences so as to react more suffi-
family configuration models and decision support, the ciently to new customers’ requirements or to adapt to
coherence among the product architecture, family, and a changing business environment. Substantial research
platform, and so on. Further issues may consider, for is needed to transfer the tools and principles from this
instance, to what extent a product family architecture area to an enterprise that is no longer based on the
and platform can best represent the capability of an manufacturing of products but on interactions with each
enterprise? How can product families be matched with individual customer. It is necessary to establish method-
an existing set of resources and enterprise capabilities? ologies for describing capabilities and sharing them in
How can various players (customers, designers, suppli- an extended value chain network. For example, how
ers, production engineers, etc.) communicate well within can modular process models be created and configured
the same platform of product family design? How to in order to integrate the capabilities of different firms
evolve product platforms and architectures in accor- in the fulfillment process of a specific customer order?
dance with changes in customers’ requirements, product The idea of inter-organizational cooperation and vir-
technologies and enterprise capabilities? How to coor- tual enterprises has to be developed much further. The
dinate basic product designs with variant design of the benefits of integrating suppliers into customized manu-
configuration process? Last but not least, what are the facturing and capability development are well described
key factors that contribute to design-by-customer with in theory, but not well implemented in practice. Further
22 J Intell Manuf (2007) 18:5–29

research is needed to establish scalable and transfer- from cost control to value creation, existing account-
able coordination schemes in the logistics domain as ing and control systems are mostly dominated by the
well. Besides, the tradeoff of delayed satisfaction and practice of product costing. Savings and additional costs
increase in volume efficiency is obvious from the manu- resulting from different degrees of interaction with the
facturer perspective, but it may not be the case for cus- customers are not covered by most industrial accounting
tomers. This leads to the justification of the congruence systems. Activity-based costing and the balanced score
of customer specifications and manufacturing capabili- card approaches may provide initial solutions; however,
ties, for example, to explain how major ERP suites (e.g. approved ratios for calculating the value of customer
SAP, Baan, J.D. Edwards, etc.) affect operational perfor- relationships are still missing; nor are parameters for
mances when implementing product families (Medina & evaluating the extent of the market research informa-
Duffy, 1998). tion gained by aggregated customer knowledge. More-
over, the value contribution of product families should
Economic justification be evaluated from the customers’ perspective. There is
rarely any attempt to explicitly measure the need for
Product family design and development is associated individualization or to quantify the value of product
with new cost and profit structures that can be coined as families from a customers’ perspective. The issue of jus-
“economies of scale and scope”. Current research on the tifying the economic value of introducing individualized
economic and performance evaluation of product fami- products is of vital importance. Only if the increment in
lies is dominated by empirical studies, ad hoc samples, or the customer-perceived value or utility suffices enough
broad approaches based on cost accounting. Traditional can product customization become a mass phenome-
cost accounting by allocating fixed costs and variable non. Recent study on the valuation of flexibility has
costs across multiple products may produce distorted suggested that the real option approach surmounts tra-
cost-carrying figures due to possible sunk costs associ- ditional DCF analysis-based methods that tend to ignore
ated with investment into product and process platforms the upside potentials from management flexibility (Jiao
(Jiao & Zhang, 2005). It is quite common in product et al., 2005a; Otto et al., 2003).
family fulfillment that design and manufacturing admit Furthermore, the risks related to product family deve-
resources, and thus the related costs, to be shared among lopment need to be addressed properly. Robertson and
multiple products in a reconfigurable fashion, as well as Ulrich (1998) observe the organizational risks related to
per-product fixed costs (Moore et al., 1999). Yano and platform development. Developing product platforms in
Dobson (1998) observe a number of industrial settings, most cases requires more investments and development
where a wide range of products are produced with very time than developing a single product, which may delay
little incremental costs per se, or very high develop- the time to market and affect the return on investment
ment costs are shared across broad product families, or time. Meyer and Lehnerd (1997) point out such risks that
fixed costs and variable costs change dramatically with weak common platform may undermine the competi-
product variety. They point out that “the accounting sys- tiveness of the entire product line and therefore a broad
tems, whether traditional or activity-based, do not sup- array of products may feel the pain. In addition to fixed
port the separation of various cost elements”. Safizadeh, investments, developing platforms may result in over-
Ritzman, and Mallick (2000) derive similar results from design of low-end product variants in order to enable
an empirical study of 142 manufacturing plants: plants reuse with high-end products (Krishnan & Gupta, 2001).
that provide a high degree of customization incurs high Henderson and Clark (1990) identify one potential neg-
cost structures. However, when controlling for produc- ative effect of modular product architectures that origi-
tion processes the tradeoff disappears. This means once nates from the risk of creating barriers to architectural
a company has defined its product range along with an innovation. Organizational forces may also hinder the
appropriate production process, product family-based ability to balance commonality and distinctiveness (Hal-
customization that falls within the range offered does man, Hofer, & van Wuuren, 2003).
not cost any extra.
The economic justification of product families req- Extended platforms for collaborative product families
uires the identification of proper measures and perfor-
mance indicators to characterize different outcomes of Product family design and development enhances profit-
a product customization system. This task is imperative ability through a synergy of increased customer-per-
because the current accounting systems are not designed ceived value and cost reduction in design, manufacturing
for assessing the true economical benefits from the total and the supply chain. In line with the holistic view shown
value chain point of view. Even if the focus is shifted in Figure 1, a product family should ideally be built on
J Intell Manuf (2007) 18:5–29 23

sharing a multidimensional core of assets such as stan- manufacturing planning, and supply chain management
dardized components, manufacturing, supply and dis- within a coherent framework.
tribution processes, customer segmentation and brand While the field of product families has matured rap-
positioning (Jiao, Huang, and Tseng, 2004). To sup- idly over the last decade, there are still a number of rela-
port coordination of the demand and supply chains with tively unexplored topics that offers numerous opportu-
product families, it is necessary to extend platform think- nities for scholarly inquiry. Despite the diligent efforts
ing to the entire continuum of product fulfillment, inclu- of many scholars, much more research needs to be put
ding customer platforms, brand platforms, product plat- into the agenda. As discussed in this review, the unan-
forms, process platforms, and global platforms (Halman swered questions may be examined through a wide vari-
et al., 2003). Greater complexity must be introduced to ety of approaches, both theoretically and methodologi-
product family design decisions when considering more cally. We hope that the review will serve as a basic frame
decision variables or parameters pertinent to the coor- of reference and will spur more research to examine this
dination across the product, manufacturing process and important and fascinating field.
supply chain domains (Rungtusanatham & Forza, 2005).
It is also important to incorporate the pervasive con- Acknowledgements The authors would like to acknowledge sup-
port from the National Science Foundation under Career Award
nectivity of the Internet to coordinate the participation DMI-0133923. Any opinions, findings, and conclusions or recom-
of all parties including customers, suppliers, service pro- mendations presented in this paper are those of the authors and
viders and many others. While the Internet facilitates do not necessarily reflect the views of the National Science Foun-
a company’s communication with its customers to con- dation.
figure its products and even offer online transactions,
the manufacturers must implement such web-based
solutions to allow them to interactively communicate
information related to product design, development, References
manufacturing, and logistics within their own infrastruc-
tures in a coherent manner. The ultimate goal is to Agard, B., & Kusiak, A. (2004). Data-mining-based methodol-
achieve an F2B2C e-Commerce that aims to realize ogy for the design of product families. International Journal of
Production Research, 42(15), 2955–2969.
product customization over the Internet through coher-
Ahmad, S., & Schroeder, R.G. (2002). Refining the
ently integrating manufacturing production automation product-process matrix. International Journal of Opera-
with supply chain management and sales-service sup- tions & Production Management, 22(1), 103–124.
port into a collaborative web of interactive commerce Akundi, S., Simpson, T. W., & Reed, P. M. (2005). Multi-objective
design optimization for product platform and product family
(Lu, 2004).
design using genetic algorithms. In ASME design engineer-
ing technical conferences, DETC2005/DAC-84905, Long Beach,
CA.
Allen, K. R., & Carlson-Skalak, S. (1998). Defining product archi-
tecture during conceptual design. In ASME design engineering
Summary technical conference, DETC98/DTM-5650, Atlanta, GA.
Appelqvist, P., & Heikkilä, J. (2003). A model for product, pro-
Substantial progress has been achieved in the areas of cess and supply chain configuration in a build-to-order environ-
product family design optimization, product family con- ment. The 4th Doctoral Colloquium in Industrial Engineering
and Management, Espoo, Finland.
figuration, modular architectures, and product portfolio Ardissono, L., Felfernig, A., Friedrich, G., Goy, A., Jannach, D.,
planning. Future research lies in the holistic view and Petrone, G., Schäfer, R., & Zanker, M. (2003). A framework
system-wide solutions. More specifically, product family for the development of personalized, distributed web-based
design needs to incorporate more front-end issues such configuration systems. AI Magazine, 24(3), 93–110.
Aydiny, A. O., & Gugor, A. (2005). Effective relational database
as explicit customer modeling and integration, product approach to represent bills-of-materials. International Journal
demand and market segmentation, along with the eco- of Production Research, 43(6), 1143–1170.
nomic evaluation of product families. In the meantime, Baker, K. R., Magazine, M. J., & Nuttle, H. L. W. (1986). The effect
the spectrum of product family design should extend of commonality on safety stock in a simple inventory model.
Management Science, 32(8), 983–988.
to include more back-end issues involving manufactur- Balakrishnan, P. V. S., & Jacob, V. S. (1996). Genetic algorithms
ing, production, and the supply chain. The fulfillment for product design. Management Science, 42(1), 1105–1117.
of product families requires alignment of the customer, Baldwin, C. Y., & Clark, K. B. (2000). Design rules: The power of
product, process and supply chain decisions. Extended modularity. Cambridge, MA: MIT Press.
Banerjee, P., & de Weck, O. L. (2004). Flexibility strategy – valuing
configure-to-order platforms are the core to support flexible product options. In INCOSE/ICSE conference on syn-
product customization over the Internet while achiev- ergy between systems engineering and project management, Las
ing a synergy of sales force automation, product design, Vegas, Nevada.
24 J Intell Manuf (2007) 18:5–29

Bei, Y., & MacCallum, K. J. (1995). Decision support for con- of configurable components. In ASME design engineering tech-
figuration management in product development. In The 3rd nical conferences, DETC/DTM-21714, Pittsburgh, PA.
international conference on computer integrated manufacturing Collier, D. A. (1981). The measurement and operating bene-
(Vol. 1, pp. 278–286). Singapore: World Scientific. fits of component part commonality. Decision Sciences, 12(1),
Benjaafar, S., Heragu, S. S., & Irani, S. A. (2002). Next genera- 85–96.
tion factory layouts: Research challenges and recent progress, Corbett, B., & Rosen, D. W. (2004). A configuration design based
Interfaces, 32(6), 58–76. method for platform commonization for product families. AIE-
Bi, Z. M., & Zhang, W. J. (2001). Modularity technology in DAM, 18(1), 21–39.
manufacturing: Taxonomy and issues. International Journal of Costa, C. A., & Young, R. I. M. (2001). Product range models sup-
Advanced Manufacturing Technology, 18(5), 381–390. porting design knowledge reuse. Proceedings of IMechE, Part
Blackenfelt, M. (2000a). Profit maximisation while considering B, Journal of Engineering Manufacture, 215(3), 323–337.
uncertainty by balancing commonality and variety using robust Cutherell, D. (1996). Product architecture. In M. Rosenau, G. Grif-
design – the redesign of a family of lift tables. In ASME fin, G. Castellion & N. Anschuetz (Eds.), The PDMA handbook
design engineering technical conferences, DETC2000/DFM- of new product development. John Wiley & sons.
14013, Baltimore, MD. Dahmus, J. B., & Otto, K. N. (2001). Incorporating lifecycle costs
Blackenfelt, M. (2000b). Design of robust interfaces in modular into product architecture decisions. In ASME design engineer-
products. In ASME design engineering technical conferences, ing technical conferences, DETC2001/DAC-21110, Pittsburgh,
DETC00/DAC-14486, Baltimore, MD. PA, ASME.
Blackhurst, J., Wu, W., & O’Grady, P. (2005). PCDM: A decision Dai, Z., & Scott, M. J. (2004). Product platform design through sen-
support modeling methodology for supply chain, product and sitivity analysis and cluster analysis. In ASME design engineer-
process design decisions. Journal of Operations Management, ing technical conferences, DETC2004/DAC-57464, Salt Lake
23(3–4), 325–343. City, UT.
Blecker, T., Abdelkafi, N., Kreuter, G., & Friedrich, G. (2004a). De Lit, P. G., & Delchambre, A. (2003). Integrated design of a
Product configuration systems: State-of-the-art, conceptualiza- product family and its assembly system. Massachusetts: Kluwer
tion and extensions. In A. B. Hamadou, F. Gargouri, & M. Academic Publishers.
Jmaiel (Eds.), Intelligence Articielle (MCSEAI 2004). Tunisia: De Weck, O. L., Suh, E. S., & Chang, D. (2003). Product family and
Génie logiciel & Soussee. platform portfolio optimization. In ASME design engineering
Blecker, T., Abdelkafi, N., Kreutler, G., & Friedrich, G. (2004b). technical conferences, DETC03/DAC-48721, Chicago, IL.
An advisory system for customers’ objective needs elicitation Dobrescu, G., & Reich, Y. (2003). Progressive sharing of mod-
in mass customization. In The 4th international ICSC sympo- ules among product variants. Computer-Aided Design, 35(9),
sium on engineering of intelligent systems. Funchal, Portugal: 791–806.
University of Madeira. Dobson, G., & Yano, C.A. (1994). Product line and tech-
Bohm, M. R., & Stone, R. B. (2004). Representing functional- nology selection with shared manufacturing and engineer-
ity to support reuse: Conceptual and supporting functions. In ing design resources, Paper provided by Rochester, Busi-
ASME design engineering technical conferences, DETC2004- ness - Center for Manufacturing and Operations Management
57693, Salt Lake City, Utah. in its paper series with number (RePEc:fth:robuma:95-01),
Bozarth, C., & McDermott, C. (1998). Configurations in manu- http://ideas.repec.org/p/fth/robuma/95-01.html.
facturing strategy: a review and directions for future research, Doran, D. (2003). Supply chain implications of modularization.
Journal of Operations Management, 16(3), 427–439. International Journal of Operations & Production Management,
Brabazon, P. G., & MacCarthy, B. (2004). Virtual-build-to-order 23(3/4), 316–326.
as a mass customization order fulfilment model. Concurrent Du, X., Jiao, J., & Tseng, M. M. (2001). Architecture of product
Engineering: Research and Application, 12(2), 155–165. family: Fundamentals and methodology. Concurrent Engineer-
Bramham, J., & MacCarthy, B. (2003). Matching configurator attri- ing: Research and Application, 9(4), 309–325.
butes to business strategy. In The 2nd world congress on mass Du, X., Jiao, J., & Tseng, M. M. (2002a). Graph grammar based
customization and personalization, Munich. product family modeling. Concurrent Engineering: Research
Chakravarty, A. K., & Balakrishnan, N. (2001). Achieving prod- and Application, 10(2), 113–128.
uct variety through optimal choice of module variations. IIE Du, X., Jiao, J., & Tseng, M. M. (2002b). Product family model-
Transactions, 33(7), 587–598. ing and design support: An approach based on graph rewriting
Chandrasekaran, B., Stone, R. B., & Mcadams, D. A. (2004). systems. AIEDAM, 16(2), 103–119.
Developing design templates for product platform focused Eppinger, S. D., Whitney, D. E., Smith, R. P., & Gebala, D. A.
design. Journal of Engineering Design, 15(3), 209–228. (1994). A model-based method for organizing tasks in product
Chen, L., & Lin, L. (2002). Optimization of product configuration development. Research in Engineering Design, 6(1), 1–13.
design using functional requirements and constraints. Research Erens, F., & Verhulst, K. (1997). Architectures for product fami-
in Engineering Design, 13(3), 167–182. lies. Computers in Industry, 33(2–3), 165–178.
Choi, S. C., Desarbo, W. S., & Harker, P. T. (1990). Product posi- Ericsson, A., & Erixon, G. (1999). Controlling Design Variants:
tioning under price competition. Management Science, 36(2), Modular Product Platforms, New York: ASME.
175–199. Erixon, G., & Ostgren, B. (1993). Synthesis and evaluation tool
Choi, T. Y., & Hong, Y. (2002). Unveiling the structure of supply for modular designs. In International conference on engineering
networks: Case studies in Honda, Acura, and DiamlerChrysler. design ( pp. 898–905). Hague, Netherlands.
Journal of Operations Management, 20(5), 469–493. Erlandsson, A., Erixon, G., & Ostgren, B. (1992). Product modules
Chong, J. K., Ho, T. H., & Tang, C. S. (1998). Product structure, – the link between QFD and DFA? The International Forum on
brand width and brand share. In T. H. Ho & C. S. Tang (Eds.), Product Design for Manufacture and Assembly, Newport, RI.
Product variety management: Research advances, USA: Kluwer Farrell, R., & Simpson, T. W. (2001). Improving commonal-
Academic Publisher. ity in custom products using product families. In ASME
Claesson, A., Johannesson, H., & Gedell, S. (2001). Platform prod- design engineering technical conferences, DETC2001/DAC-
uct development: Product model – a system structure composed 21125, Pittsburgh, PA.
J Intell Manuf (2007) 18:5–29 25

Feitzinger, E., & Lee, H. L. (1997). Mass customization at Hew- Gershenson, J. K., Prasad, G. J., & Zhang, Y. (2003). Product mod-
lett-Packard: The power of postponement. Harvard Business ularity: Definitions and benefits. Journal of Engineering Design,
Review, 75(1), 116–121. 14(3), 295–313.
Felfernig, A., Friedrich, G., & Jannach, D. (2001). Conceptual Gonzalez-Zugasti, J. P., Otto, K., & Baker, J. (2000). A method
modeling for configuration of mass customizable products. Arti- for architecting product platforms. Research in Engineering
ficial Intelligence in Engineering, 15(2), 165–176. Design, 12(2), 61–72.
Fellini, R., Kokkolaras, M., Papalambros, P. Y., & Perez-Duarte, A. Green, P. E., & Krieger, A. M. (1985). Models and heuristics for
(2002a). Platform selection under performance loss constraints product line selection. Marketing Science, 4(1), 1–19.
in optimal design of product families. In ASME design engi- Guo, F., & Gershenson, J. K. (2004). A comparison of modu-
neering technical conferences, DETC02/DAC-34099, Montreal, lar product design methods on improvement and iteration. In
Canada. ASME design engineering technical conferences, Salt Lake City,
Fellini, R., Kokkolaras, M., Michelena, N., Papalambros, P., Saitou, UT.
K., Perez-Duarte, A., & Fenyes, P. (2002b). A sensitivity-based Gupta, S., & Krishnan, V. (1998). Product family-based assem-
commonality strategy for family products of mild variation, bly sequence design methodology. IIE Transactions, 30(10),
with application to automotive body structures. In The 9th 933–945.
AIAA/ISSMO symposium on multidisciplinary analysis and Gupta, S., & Krishnan, V. (1999). Integrated component and sup-
optimization, AIAA-2002-5610, Atlanta, GA. plier selection for a product family. Production and Operations
Finch, W. (1999). Set-based models of product platform design and Management, 8(2), 163–182.
manufacturing processes. In ASME design engineering techni- Halman, J. I. M., Hofer, A. P., & van Wuuren, W. (2003). Platform-
cal conferences, DETC99/DTM-8763, Las Vegas, NV. driven development of product families: Linking theory with
Fine, C. H., Golany, B., & Naseraldin, H. (2005). Modeling trade- practice. Journal of Product Innovation Management, 20(2),
offs in three-dimensional concurrent engineering: A goal pro- 149–162.
gramming approach. Journal of Operations Management, 23 Hayes, R., & Wheelwright, S. (1984). Restoring our Competitive
(3–4), 389–403. Edge. New York, NY: Wiley.
Fisher, M., Ramdas, K., & Ulrich, K. (1999). Component sharing Hazelrigg, G. A. (1998). A framework for decision-based engi-
in the management of product variety: A study of automotive neering design. ASME Journal of Mechanical Design, 120(4),
braking systems. Management Science, 45(3), 297–315. 653–658.
Fixson, S. K. (2002). Linking modularity and cost: A methodology He, D. W., & Kusiak, A. (1996). Performance analysis of modular
to assess cost implications of product architecture differences to products. International Journal of Production Research 34(1),
support product design, Ph.D. Thesis, Massachusetts Institute 253–272.
of Technology, Cambridge, MA. Hegge, H. M. H., & Wortmann, J. C. (1991). Generic bill-of-mate-
Fixson, S. K. (2005). Product architecture assessment: A tool to rial: A new product model. International Journal of Production
link product, process, and supply chain design decisions. Journal Economics, 23(1–3), 117–128.
of Operations Management, 23(3–4), 345–369. Henderson, R. M., & Clark, K. B. (1990). Architecture innova-
Forza, C., Salvador, F. (2002). Managing for variety in the order tion: The reconfiguration of existing product technologies and
acquisition and fulfillment process: The contribution of prod- the failure of established firms. Administrative Science Quar-
uct configuration systems. International Journal of Production terly, 35(1), 9–30.
Economics, 76(1), 87–98. Hernandez, G., Allen, J. K., Woodruff, G. W., Simpson, T. W.,
Fredriksson, P. (2002). Modular assembly in the car industry – Bascaran, E., Avila, L. F., & Salinas, F. (2001). Robust design of
an analysis of organizational forms’ influence on performance. families of products with production modeling and evaluation.
European Journal of Purchasing & Supply Management, 8(4), ASME Journal of Mechanical Design, 123(2), 183–190.
221–233. Hill, T. (1994). Manufacturing strategy: Text and cases. Homewood,
Fujimoto, H., & Ahmed, A. (2003). Assembly process design for IL: Irwin.
managing manufacturing complexities because of product vari- Ho, T. H., & Tang, C. S. (1998). Product variety management:
eties. International Journal of Flexible Manufacturing Systems, Research advances. Boston: Kluwer Academic Publishers.
15(4), 283–307. Hölttä, K., & Salonen, M. (2003). Comparing three modularity
Fujita, K., Sakaguchi, H., & Akagi, S. (1999). Product vari- methods. In ASME design engineering technical conferences,
ety deployment and its optimization under modular archi- DETC2003/DTM-48649, Chicago, IL.
tecture and module commonalization. In ASME design engi- Hölttä, K., Tang, V., & Seering, W. (2003). Modularizing product
neering technical conferences, DETC99/DFM-8923, Las Vegas, architectures using dendrograms. In International conference
NV. on engineering design, Stockholm.
Fujita, K., & Yoshida, H. (2004). Product variety optimization Huang, C.-C. (2004). A multi-agent approach to collaborative
simultaneously designing module combination and module design of modular products. Concurrent Engineering: Research
attributes. Concurrent Engineering: Research and Application, and Application, 12(1), 39–47.
12(2), 105–118. Huang, C.-C., & Liang, W. Y. (2001). A formalism for design-
Garg, A. (1999). An application of designing products and pro- ing with modules. Journal of the Chinese Institute of Industrial
cesses for supply chain management. IIE Transactions, 31(5), Engineers, 18(3), 13–20.
417–429. Huang, G. Q., Zhang, X. Y., & Liang, L. (2005). Towards integrated
Georgiopoulos, P., Fellini, R., Sasena, M., & Papalambros, P. optimal configuration of platform products, manufacturing pro-
Y. (2002). Optimal design decisions in product portfolio val- cesses, and supply chains. Journal of Operations Management,
uation. In ASME design engineering technical conferences, 23(3–4), 267–290.
DETC2002/DAC-34097, Montreal, Canada. Huffman, C., & Kahn, B. (1998). Variety for sale: Mass customiza-
Gerchak, Y., Magazine, M. J., & Gamble, A. B. (1988). Compo- tion or mass confusion? Journal of Retailing, 74(4), 491–513.
nent commonality with service level requirements. Manage- Hvam, L. (2004). A multi-perspective approach for the design
ment Science, 34(6), 753–760. of product configuration systems – an evaluation of industry
26 J Intell Manuf (2007) 18:5–29

applications. In International conference on economic, techni- extension. Journal of Product Innovation Management, 18(4),
cal and organizational aspects of product configuration systems. 219–230.
Lyngby, Denmark: Technical University of Denmark. Klocke, F., Fallbohmer, M., Kopner, A., & Trommer, G.
Jiao, R. J., Huang, G. G. Q., & Tseng, M. M. (2004). Concurrent (2000). Methods and tools supporting modular process design.
enterprising for mass customization. Concurrent Engineering: Robotics and Computer Integrated Manufacturing, 16(13),
Research and Application, 12(2), 83–88. 411–423.
Jiao, J., Lim, C. M., & Kumar, A. (2005a). Product family con- Kolisch, R. (2000). Integration of assembly and fabrication for
figuration design based on hybrid real options valuation. IIE make-to-order production. International Journal of Production
Transactions. Economics, 68(3), 287–306.
Jiao, J., Pokharel, S., Zhang, L., & Zhang, Y. (2005b). Coordina- Kogut, B., & Kulatilaka, N. (1994). Option thinking and platform
tion of product and process variety in mass customization with investment: Investing in opportunity. California Management
data mining approach. In The 10th annual international confer- Review, 36(20), 52–71.
ence on industrial engineering theory, applications & practice, Kokkolaras, M., Fellini, R., Kim, H. M., Michelena, N. F., &
Clearwater Beach, FL. Papalambros, P. Y. (2002). Extension of the target cascading
Jiao, J., & Tseng, M. M. (1999). A methodology of developing formulation to the design of product families. Structural and
product family architecture for mass customization. Journal of Multidisciplinary Optimization, 24(4), 293–301.
Intelligent Manufacturing, 10(1), 3–20. Kota, S., Sethuraman, K., & Miller, R. (2000). A metric for evaluat-
Jiao, J., & Tseng, M. M. (2000). Understanding product family for ing design commonality in product families. Journal of Mechan-
mass customization by developing commonality indices. Jour- ical Design, 122(4), 403–410.
nal of Engineering Design, 11(3), 225–243. Krishnan, V., & Gupta, S. (2001). Appropriateness and impact
Jiao, J., & Tseng, M. M. (2004). Customizability analysis in of platform-based product development. Management Science,
design for mass customization. Computer-Aided Design, 36(8), 47(1), 52–68.
745–757. Kurtadikar, R. M., & Stone, R. B. (2003). Investigation of cus-
Jiao, J., Tseng, M. M., Duffy, V. G., & Lin, F. (1998). Product fam- tomer needs frequency vs. weight in product platform plan-
ily modeling for mass customization. Computers and Industrial ning. In ASME international mechanical engineering congress
Engineering, 35(3–4), 495–498. and R&D Expo, IMECE2003-42786, Washington, DC.
Jiao, J., Tseng, M. M., Ma, Q., & Zou, Y. (2000). Generic bill of Kusiak, A. (2002). Integrated product and process design: A
materials and operations for high-variety production manage- modularity perspective. Journal of Engineering Design, 13(3),
ment. Concurrent Engineering: Research and Application, 8(4), 223–231.
297–322. Kusiak, A., & Huang, C. C. (1996). Development of modular prod-
Jiao, J., Zhang, L., & Pokharel, S. (2003). Process platform plan- ucts. IEEE Transactions on Components, Packaging, and Man-
ning for mass customization. In The 2nd interdisciplinary world ufacturing Technology. Part A, 19(4), 523–538.
congress on mass customization and personalization, Munich, Lancaster, K. (1990). The economics of product variety: A survey.
Germany. Marketing Science, 9(3), 189–211.
Jiao, J., Zhang, L., & Pokharel, S. (2005c). Coordinating prod- Lee, H., & Billington, C. (1994). Designing products and pro-
uct and process variety for mass customized order fulfillment. cesses for postponement. In S. Dasu & C. Eastmen (Eds.), Man-
Production Planning and Control, 16(6), 608–620. agement of design: engineering and management perspectives.
Jiao, J., Zhang, L., & Prasanna, K. (2005d). Process variety mod- Boston: Kluwer Academic Publishers.
eling for process configuration in mass customization: An Lee, H. L., & Sasser, M. M. (1995). Product universality and design
approach based on object-oriented Petri-nets with change- for supply chain management. Production Planning & Control,
able structures. International Journal of Flexible Manufacturing 6(3), 270–277.
Systems, 16, 335–362. Lee, H. L., & Tang, C. S. (1998). Variability reduction through
Jiao, J., & Zhang, Y. (2005). Product portfolio planning with operations reversal. Management Science, 44(2), 162–172.
customer–engineering interaction. IIE Transactions, 37(9), Li, H., & Azarm, S. (2002). An approach for product line design
801–814. selection under uncertainty and competition. Transactions of
Joglekar, N., & Rosenthal, R. (2003). Coordination of design sup- the ASME, Journal of Mechanical Design, 124(3), 385–392.
ply chains for bundling physical and software products. Journal Liang, W. Y., & O’Grady, P. (2000). A constrained evolutionary
of Product Innovation Management, 20(5), 374–390. search formalism for remote design with modules. International
Johannesson, H. (1992). Computer aided modeling of families and Journal of Computer Integrated Manufacturing, 13(2), 65–79.
family members of designed parts. Advanced Design Automa- Lu, S. C.-Y. (2004). F2B2C e-Commerce, http://www.f2b2c.com/.
tion, 44(2), 173–179. Malmström, J., & Malmqvist, J. (1998). Tradeoff analysis in prod-
Johannesson, H., & Claesson, A. (2005). Systematic product plat- uct structures: A case study at Celsius Aerotech. Proceedings
form design: A combined function-means and parametric mod- of NordDesign’98 (pp. 187–196). Stockholm.
eling approach. Journal of Engineering Design, 16(1), 25–43. Männistö, T. (2000). A conceptual modeling approach to product
Kamrani, A. K., & Gonzalez, R. (2003). A genetic algorithm-based families and their evolution. Ph.D thesis, Helsinki University
solution methodology for modular design. Journal of Intelligent of Technology, Acta Polytechnica Scandinavica, Mathematics
Manufacturing, 14(6), 599–616. and Computing Series, No. 106, Espoo.
Kaul, A., & Rao, V. R. (1995). Research for product position- Markus, A., & Váncza, J. (1998). Product line development with
ing and design decisions: An integrative review. International customer interaction. CIRP Annals, 47(1), 361–364.
Journal of Research in Marketing, 12(4), 293–320. Martin, M. V., & Ishii, K. (1997). Design for variety: Development
Kim, B., Leung, J. M. Y., Park, K. T., Zhang, G., & Lee, S. (2002). of complexity indices and design charts. In ASME design engi-
Configuring a manufacturing firm’s supply network with multi- neering technical conferences, DFM-4359, Sacramento, CA.
ple suppliers. IIE Transactions, 34(8), 663–677. Martin, M. V., & Ishii, K. (2002). Design for variety: Developing
Kim, K., & Chhajed, D. (2001). An experimental investigation of standardized and modularized product platform architectures.
valuation change due to commonality in vertical product line Research in Engineering Design, 13(4), 213–235.
J Intell Manuf (2007) 18:5–29 27

Maull, R., Hughes, D., & Bennett, J. (1992). The role of the bill-of- Momme, J., Moeller, M. M., & Hvolby, H.-H. (2000). Linking
materials as a CAD/CAPM interface and the key importance of modular product architecture to the strategic sourcing process:
engineering change control. Computing & Control Engineering Case studies of two Danish industrial enterprises. International
Journal, 3(2), 63–70. Journal of Logistics: Research and Applications, 3(2), 127–146.
Martinez, M.T., Favrel, J., & Fhodous, P. (2000). Product family Moore, W. L., Louviere, J. J., & Verma, R. (1999). Using conjoint
manufacturing plan generation and classification. Concurrent analysis to help design product platforms. Journal of Product
Engineering: Research and Applications, 8(1), 12–22. Innovation Management, 16(1), 27–39.
Maupin, A. J., & Stauffer, L. A. (2000). A design tool to help Morgan, L. O., Daniels, R. L., & Kouvelis, P. (2001). Market-
small manufacturers reengineer a product family. In ASME ing/Manufacturing tradeoffs in product line management. IIE
design engineering technical conferences, DETC99/DTM- Transactions, 33(11), 949–962.
14568, Baltimore, MD. Muffatto, M., & Roveda, M. (2002). Product architecture and plat-
McAdams, D. A., & Wood, K. L. (2002). A quantitative similar- forms: A conceptual framework. International Journal of Tech-
ity metric for design-by-analogy. ASME Journal of Mechanical nology Management, 24(1), 1–16.
Design, 124(2), 173–182. Nayak, R. U., Chen, W., & Simpson, T. W. (2002). A variation-
McCarthy, I., & Anagnostou, A. (2004). The impact of outsourc- based methodology for product family design. Journal of Engi-
ing on the transaction costs and boundaries of manufacturing. neering Optimization, 34(1), 65–81.
International, Journal of Production Economics, 88(1), 61–71. Nelson, S. A. II, Parkinson, M. B., Papalambros, P. Y. (2001).
McGrath, M. (1995). Product strategy for high-technology compa- Multicriteria optimization in product platform design. ASME
nies. New York: Irwin Professional Publishing. Journal of Mechanical Design, 123(2), 199–204.
McIvor, R., & Humphreys, P. (2004). Early supplier involvement Newcomb, P. J., Bras, B., & Rosen, D. W. (1996). Implications
in the design process: Lessons from the electronics industry. of modularity on product design for the life cycle. In ASME
OMEGA, 32(3), 179–199. design engineering technical conferences, DETC96/DTM-1516,
McKay, A., Erens, F., & Bloor, M. S. (1996). Relating product Irvine, CA.
definition and product variety. Research in Engineering Design, Ng, N. K., & Jiao, J. (2004). A domain-based reference model
8(2), 63–80. for the conceptualization of factory loading allocation prob-
McKay, A., & Pennington, A. D. (2001). Towards an integrated lem in multi-site manufacturing supply chains. Technovation,
description of product, process and supply chain. International 24(8), 631–642.
Journal of Technology Management, 21(3/4), 203–220. Nobelius, D., & Sundgren, N. (2002). Managerial issues in parts
Medina, J. F., & Duffy, M. F. (1998). Standardization vs globaliza- sharing among product development projects: A case study.
tion: A new perspective of brand strategies. Journal of Product Journal of Engineering Technology Management, 19(1), 59–73.
and Brand Management, 7(3), 223–243. Novak, S., & Eppinger, S. D. (2001). Sourcing by design: Product
Mehrabi, M. G., Ulsoy, A. G., & Koren, Y. (2000). Reconfigurable architecture and the supply chain. Management Science, 47(1),
manufacturing systems: Key to future manufacturing. Journal 189–204.
of Intelligent Manufacturing, 11(4), 413–419. O’Grady, P., & Liang, W.-Y. (1998). An object oriented approach
Messac, A., Martinez, M. P., & Simpson, T. W. (2002a). Effec- to design with modules. Computer Integrated Manufacturing
tive product family design using physical programming and Systems, 11(4), 267–283.
the product platform concept exploration method. Engineer- Olsen, K. A., & Stre, P. (1996). Managing product variability by
ing Optimization, 3(3), 245–261. virtual products. International Journal of Production Research,
Messac, A., Martinez, M. P., & Simpson, T. W. (2002b). A penalty 35(8), 2093–2107.
function for product family design using physical programming. Oosterman, B. (2001). Improving product development projects by
ASME Journal of Mechanical Design, 124(2), 164–172. matching product architecture and organization. Ph.D. Disser-
Meyer, M. H., & Dalal, D. (2001). Managing platform architec- tation, University of Groningen, Groningen, the Netherlands.
tures and manufacturing processes for nonassembled products. Oosterman, B., Land, M. L., & Gaalman, G. (2000). The influ-
Journal of Product Innovation Management, 19(4), 277–293. ence of shop characteristics on workload control. International
Meyer, M., & Lehnerd, A. P. (1997). The power of product plat- Journal of Production Economics, 68(1), 107–119.
form – building value and cost leadship. New York: Free Press. Otto, K., Gonzalez-Zugasti, J., & Dahmus, J. (2000). Modular
Meyer, M., & Utterback, J. (1993). The product family and the product architecture. In ASME design engineering technical
dynamics of core capability. Sloan Management Review, Spring conferences, DETC2000/DTM-4565, Baltimore, MD.
1993, 29–47. Otto, K., Tang, V., & Seering, W. (2003). Establishing quantitative
Michalek, J. J., Feinberg, F. M., & Papalambros, P. Y. (2005). economic value for features and functionality of new prod-
Linking marketing and engineering product design decisions ucts and new services, Chapter N, MIT PDMA Toolbook II,
via analytical target cascading. Journal of Product Innovation http://hdl.handle.net/1721.1/3821.
Management, 22(1), 42–62. Pagh, J. D., & Cooper, M.C. (1998) Supply chain postponement
Mikkola, J. H., & Gassmann, O. (2003). Managing modularity and speculation strategies: how to choose the right strategy,
of product architectures: Towards an integrated theory. IEEE Journal of Business Logistics, 19(2), 13–33.
Transactions on Engineering Management, 50(2), 204–218. Park, B., Ghosh, S., & Murthy, N. N. (2000). A framework for
Mikkola, J. H., & Skjøtt-Larsen, T. (2004). Supply-chain inte- integrating product platform development with global supply
gration – implications for mass customization, modularization chain configuration. In National DSI conference, Orlando, FL.
and postponement strategies. Production Planning & Control, Park, J., & Simpson, T. W. (2005). Development of a production
15(4), 352–361. cost estimation framework to support product family design.
Molina, A., Rodriguez, C. A., Ahuett, H., Cortes, J. A., Ra- International Journal of Production Research, 43(4), 731–772.
mirez, M., Jiménez, G., & Martinez, S. (2005). Next-genera- Perera, H. S. C., Nagarur, N., & Tabucanon, M. T. (1999). Compo-
tion manufacturing systems: key research issues in developing nent part standardization: A way to reduce the life-cycle cost
and integrating reconfigurable and intelligent machines. Inter- of products. International Journal of Production Economics,
national Journal of Computer Integrated Manufacturing, 18(7), 60–61(1), 109–116.
525–536.
28 J Intell Manuf (2007) 18:5–29

Petersen, K. J., Handfield, R. B., & Ragatz, G. L. (2005). Sup- Seepersad, C. C., Mistree, F., & Allen, J. K. (2002). A quanti-
plier integration into new product development: Coordinating tative approach for designing multiple product platforms for
product, process and supply chain design. Journal of Operations an evolving portfolio of products. In ASME design engineering
Management, 23(3–4), 371–388. technical conferences, DETC2002/DAC-34096, Montreal, Que-
Pine, B. J. (1993). Mass customization: The new frontier in business bec, Canada.
competition. Boston: Harvard Business School Press. Sharman, D. M., & Yassine, A. A. (2004). Characterizing complex
Prasad, B. (1998). Designing products for variety and how to man- product architectures. Systems Engineering, 7(1), 35–60.
age complexity. Journal of Product & Brand Management, 7(3), Schierholt, K. (2001). Process configuration: Combining the prin-
208–222. ciples of product configuration and process planning. AIE-
Raman, N., & Chhajed, D. (1995). Simultaneous determination of DAM, 15(5), 411–424.
product attributes and prices and production processes in prod- Shooter, S. B., Simpson, T. W., Kumara, S. R. T., Stone, R. B., &
uct-line design. Journal of Operations Management, 12 (3–4), Terpenny, J. P. (2005). Toward a multi-agent information infra-
187–204. structure for product family planning and mass customization.
Ramdas, K., Fisher, M., & Ulrich, K. (2003). Managing variety International Journal of Mass Customization, 1(1), 134–155.
for assembled products: Modeling component systems shar- Siddique, Z. (2001). Estimating reduction in development time
ing. Manufacturing & Service Operations Management, 5(2), for implementing a product platform approach. In ASME
142–156. design engineering technical conferences, DETC2001/CIE-
Ramdas, K., & Sawhney, M. S. (2001). A cross-functional approach 21238, Pittsburgh, PA.
to evaluating multiple line extensions for assembled products. Siddique, Z. (2005). Assembly process selection to minimize
Management Science, 47(1), 22–36. existing assembly system modification cost during new product
Robertson, D., & Ulrich, K. (1998). Planning product platforms. family member design. In ASME design engineering technical
Sloan Management Review, 39(4), 19–31. conferences, DETC2005-85016, Long Beach, CA.
Romanowski, C. J., & Nagi, R. (2002). A data mining and graph Siddique, Z., & Boddu, K. R. (2004). A mass customization infor-
theoretic approach to building generic bills of materials. In The mation framework for integration of customer in the con-
11th industrial engineering research conference, Orlando, FL. figuration/design of a customized product. AIEDAM, 18(1),
Rosen, D. W. (1996). Design of modular product architectures 71–85.
in discrete design spaces subject to life cycle issues. In ASME Siddique, Z., & Repphun, B. (2001). Estimating cost savings when
design engineering technical conferences, DETC96/DAC-1485, implementing a product platform approach. Concurrent Engi-
Irvine, CA. neering: Research and Application, 9(4), 285–294.
Rothwell, R., & Gardiner, P. (1990). Robustness and product Siddique, Z., & Rosen, D. W. (1999). Product platform design: A
design families. In M. Oakley (Ed.), Design management: A graph grammar approach. In ASME design engineering techni-
handbook of issues and methods (pp. 279–292). Cambridge, cal conferences, DETC99/DTM-8762, Las Vegas, NV.
MA: Basil Blackwell. Siddique, Z., & Rosen, D. W. (2001). On discrete design spaces for
Rungtusanatham, M., & Forza, C. (2005). Coordinating prod- the configuration design of product families. AIEDAM, 15(2),
uct design, process design, and supply chain design decisions, 91–108.
Part A: Topic motivation, performance implications, and arti- Siddique, Z., Rosen, D. W., & Wang, N. (1998). On the applicabil-
cle review process. Journal of Operations Management, 23(3–4), ity of product variety design concepts to automotive platform
257–265. commonality. In: ASME design engineering technical confer-
Sabin, D., & Weigel, R. (1998). Product configuration frameworks ences, 98-DETC/DTM-5661, Atlanta, GA.
– a survey. IEEE Intelligent Systems & their Applications, 13(4), Simpson, T. W. (2004). Product platform design and customiza-
42–49. tion: Status and promise. AIEDAM, 18(1), 3–20.
Safizadeh, M. H., Ritzman, L. P., & Mallick, D. (2000). Revisit- Simpson, T. W., D’Souza, B. (2004). Assessing variable levels of
ing alternative theoretical paradigms in manufacturing strategy. platform commonality within a product family using a multi-
Production and Operations Management, 9(2), 111–127. objective genetic algorithm. Concurrent Engineering: Research
Sako, M., & Murray, F. (2000). Modules in design, production, and Applications, 12(2), 119–130.
and use: Implications for the global automobile industry, MIT Simpson, T. W., Maier, J. R. A., & Mistree, F. (2001a). Product plat-
IMVP Annual Sponsors Meeting, Cambridge. form design: Method and application. Research in Engineering
Salhieh, S. M., & Kamrani, A. K. (1999). Macro level product Design, 13(1), 2–22.
development using design for modularity. Robotics and Com- Simpson, T. W., Nanda, J., Halbe, S., Umapathy, K., & Hodge,
puter Integrated-Manufacturing, 15(11), 319–329. B. (2003). Development of a framework for web-based prod-
Salvador, F., Forza, C., & Rungtusanatham, M. (2002). Modularity, uct platform customization. ASME Journal of Computing and
product variety, production volume, and component sourcing: Information Science in Engineering, 3(2), 119–129.
Theorizing beyond generic prescriptions. Journal of Operations Simpson, T. W., Seepersad, C. C., & Mistree, F. (2001b). Balancing
Management, 20(5), 549–575. commonality and performance within the concurrent design of
Sanchez, R. (1994). Towards a science of strategic product design: multiple products in a product family. Concurrent Engineering:
System design, component modularity, and product leverag- Research and Applications, 9(3), 177–190.
ing strategies. In The 2nd international product development Simpson, T. W., Siddique, Z., & Jiao, J. (2005). Product plat-
management conference on new approaches to development and form and product family design: Methods and applications. New
engineering, Gothenburg, Sweden. York: Springer.
Sanderson, S., & Uzumeri, M. (1995). Managing product families: Singhal, J., & Singhal, K. (2002). Supply chains and compatibility
The case of the Sony walkman. Research Policy, 24(5), 761–782. among components in product design. Journal of Operations
Sawhney, M. S. (1998). Leveraged high-variety strategies: From Management, 20(3), 289–302.
portfolio thinking to platform thinking. Journal of the Acad- Sivard, G. (2000). A generic information platform for product fam-
emy of Marketing Science, 26(1), 54–61. ilies. Doctoral thesis, Royal Institute of Technology, Stockholm.
Schwarze, S. (1996). Configuration of multiple-variant products, Sosa, M. E., Eppinger, S. D., & Rowles, C. M. (2003). The mis-
Zürich: VDF ETH. alignment of product architecture and organizational structure
J Intell Manuf (2007) 18:5–29 29

in complex product development, INSEAD Working Paper, Van Wie, M. J., Greer, J. L., Campbell, M. I., Stone, R. B.,
2003/68/TM. & Wood, K. L. (2001). Interfaces and product architec-
Stobaugh, R., & Telesio, P. (1983). Match manufacturing policies ture. In ASME design engineering technical conferences,
and product strategy. Harvard Business Review, 61(2), 113–120. DETC01/DTM-21689, Pittsburgh, PA.
Stone, R. B., Wood, K. L., & Crawford, R. H. (2000). A heuris- Van Wie, M. J., Rajan, P., Campbell, M. I., Stone, R. B., & Wood,
tic method for identifying modules for product architectures. K. L. (2003). Representing product architecture. In ASME
Design Studies, 21(1), 5–31. design engineering technical conferences, DETC2003/DTM-
Su, J. C. P., Chang, Y.-L., & Ferguson, M. (2005). Evaluation of 48668, Chicago, Illinois.
postponement structures to accommodate mass customization. Van Wie, M., Stone, R. B., Thevenot, H., & Simpson, T. (2006).
Journal of Operations Management, 23(3–4), 305–318. Examination of platform and differentiating elements in prod-
Suh, N. P. (2001). Axiomatic design: Advances and applications. uct design. Journal of Intelligent Manufacturing. Special Issue
New York: Oxford University Press. on Product Family Design and Development.
Sundgren, N. (1999). Introducing interface management in prod- Wacker, J. G., & Trelevan, M. (1986). Component part standardi-
uct family development. Journal of Production Innovation zation: An analysis of commonality sources and indices. Journal
Management, 16(1), 40–51. of Operations Management, 6(2), 219–244.
Suzue, T., & Kohdate, A. (1990). Variety reduction program: Wassenaar, H. J., & Chen, W. (2001). An approach to decision-
A production strategy for product diversification. Cambridge: based design. In Proceedings ASME 2001 design engineering
Productivity Press. technical conferences and computers and information in engi-
Takeishi, A., & Fujimoto, T. (2003). Modularization in the car neering conference, DETC2001/DTM-21683, Pittsburgh, Penn-
industry: Interlinked multiple hierarchies of product, produc- sylvania.
tion, and supplier systems. In A. Prencipe, A. Davies, & Whitney D. E. (2003). Physical limits to modularity, Working
M. Hobday (Eds.), The business of systems integration (pp. paper, ESD-WP-2003-01.03-ESD, Massachusetts Institute of
254–278). Oxford: Oxford University Press. Technology.
Tarasewich, P., & Nair, S. K. (2001). Designer-moderated product Wilhelm, B. (1997). Platform and modular concepts at Volkswa-
design. IEEE Transactions on Engineering Management, 48(2), gen – their effects on the assembly process. In K. Shimokawa, U.
175–188. Jürgens, & T. Fujimoto (Eds.), Transforming automobile assem-
Tatikonda, M. V. (1999). An empirical study of platform and deriv- bly. Berlin Heidelberg: Springer-Verlag.
ative product development projects. Journal of Product Inno- Willcox, K., & Wakayama, S. (2003). Simultaneous optimiza-
vation Management, 16(1), 3–26. tion of a multiple-aircraft family. Journal of Aircraft, 40(4),
Thevenot, H. J., Simpson, T. W. (2005). Commonality indices for 616–622.
assessing product families. In T. W. Simpson, Z. Siddique, J. Wortmann, J. C., Muntslag, D. R., & Timmermans, P. J. M. (1997).
Jiao (Eds.), Product platform and product family design: Meth- Customer driven manufacturing. London: Chapman & Hall.
ods and applications (pp. 107–129). New York: Springer. Yano, C., & Dobson, G. (1998). Profit optimizing product line
Thonemann, U. W., & Bradley, J. R. (2002). The effect of prod- design, selection and pricing with manufacturing cost consider-
uct variety on supply-chain performance. European Journal of ations. In T.-H. Ho & C. S. Tang (Eds.), Product variety man-
Operational Research, 143(3), 548–556. agement: Research advances (pp. 145–176). Kluwer Academic
Thonemann, U. W., & Brandeau, M. (2000). Optimal commonality Publisher.
in component design. Operations Research, 48(1), 1–19. Yigit, A. S., Galip-Ulsoy, A. G., & Allahverdi, A. (2002). Optimiz-
Tiihonen, J., Lehtonen, T., Soininen, T., Pulkkinen, A., Sulonen, ing modular product design for reconfigurable manufacturing.
R., & Riitahuhta, A. (1998). Modeling configurable product Journal of Intelligent Manufacturing, 13(4), 309–316.
families. The 4th WDK workshop on product structuring, Delft Yu, J. S., Gonzalez-Zugasti, J. P., & Otto, K. N. (1999). Product
University of Technology, Delft, The Netherlands. architecture definition based upon customer demands. ASME
Tseng, M. M., & Jiao, J. (1996). Design for mass customization. Journal of Mechanical Design, 121(3), 329–335.
CIRP Annals, 45(1), 153–156. Yu, T.-L., Yassine, A. A., & Goldberg, D. E. (2003). A genetic algo-
Tsubone, H., Matsuura, H., & Satoh, S. (1994). Component part rithm for developing modular product architectures. In ASME
commonality and process flexibility effects on manufacturing design engineering technical conferences, Chicago, Illinois.
performance. International Journal of Production Research, Zamirowski, E. J., & Otto, K. N. (1999). Identifying product
32(10), 2479–2493. portfolio architecture modularity using function and variety
Ulrich, K. (1995). The role of product architecture in the manu- heuristics. In ASME design engineering technical conferences,
facturing firm. Research Policy, 24(3), 419–440. DETC99/DTM-876, Las Vegas, NV.
Ulrich, K., & Eppinger, S. D. (1995). Product design and develop- Zha, X. F., Sriram, R. D., & Lu, W. F. (2004). Evaluation and
ment. New York: McGraw-Hill. selection in product design for mass customization: A knowl-
Ulrich, K., & Tung, K. (1991). Fundaments of product modularity. edge decision support approach. AIEDAM, 18(1), 87–109.
In A. Sharon (Ed.), Issues in mechanical design international Zhang, J., Wang, Q., Wan, L., & Zhong, Y. (2005). Configuration-
(pp. 73–79). New York: ASME, DE-39. oriented product modeling and knowledge management for
Van Hoek, R. I. (2001). The rediscovery of postponement: A lit- made-to-order manufacturing enterprises. International Jour-
erature review and directions for future research. Journal of nal of Advanced Manufacturing Technology, 25(1–2), 41–52.
Operations Management, 19(2), 161–184. Zinn, W., & Bowersox, D. J. (1988). Planning physical distribu-
Van Veen, E. A. (1992). Modeling product structures by generic tion with the principle of postponement. Journal of Business
bills-of-materials. New York: Elsevier. Logistics, 9(2), 117–136.

View publication stats

You might also like