Vous êtes sur la page 1sur 35

CHEMICAL PRODUCT DESIGN AND ENGINEERING

1. Chemical Product Engineering as a New Paradigm Within


Chemical Engineering

Today’s chemical industry has little in common with that of two generations ago. The
changes have been tremendous. Back in the 1960’s, the petrochemical and basic
chemistry sectors were at their height and chemical industry was focused on perhaps
50 commodity products. At that time, well-prepared chemical engineers were
expected to be thoroughly familiar with all aspects of a narrowly defined curriculum.
Nowadays, commodity plants are highly optimised, and the goals of the chemical
enterprise include not only the 50 or so commodity chemicals, but also several
thousand of what can be generically termed as engineered chemical products.
Engineered chemical products can be roughly classified into the six catego-
ries described in Table 1, encompassing single substances (like drugs) to struc-
tured formulations (such as detergents) and devices, assembled and virtual
systems. These products may have little in common based on appearance or
function, but they share the fact that they differ from commodities in three main
ways. First, their value is much higher than that of the raw materials. Second,
their commercial success is determined by performance rather than price. Finally,
their viability is highly dependent on time-to-market.
For these reasons, the design, manufacture and marketing of engineered
chemical products require a set of skills that were not crucial in the traditional
chemical industry. Flexibility, wider vision, involvement in non-traditional areas
and multidisciplinary approaches are essential to tackle the goals of the current
chemical enterprise. Chemical engineering professionals are now required to
participate in a wider variety of business decisions, and therefore this new set
of skills has become as important for success in their careers as the traditional core
skills of the discipline remain.
The need to update chemical engineering curricula in order to respond to these
changes in the chemical industry and guarantee the competitiveness of the profes-
sion has been acknowledged (1,2). The emergence of chemical product engineering
as a well-established field has been an enduring trend in this context since the early
2000’s (1,3–9). If the history of chemical engineering may be discussed traditionally
in terms of two paradigms – unit operations (developed in the 1920s and 1930s) and
transport phenomena (adopted in the late 1950s) – it is possible that chemical
product engineering becomes the discipline’s third paradigm (2,4,10).
Whilst chemical product engineering is a broad field, developing in many
directions, the simplified conceptual model in Fig. 1 may be used to elucidate its
scope and place it within the context of chemical engineering science (1).
Without precise definitions, there is often confusion between chemical
product design and chemical product engineering. However, chemical product
design is better seen as just one facet, albeit an important one, of the subject, just
as process design relates to traditional chemical engineering. Chemical product
engineering can be structured in terms of three fundamental and inter-related
pillars – (i) the chemical product pyramid; (ii) chemical product design and process
design integration; and (iii) multifaceted multiscale approach – which support
chemical product design as the major objective. Chemical product design is the

Kirk-Othmer Encyclopedia of Chemical Technology. Copyright John Wiley & Sons, Inc. All rights reserved.
Table 1. The Goals of Today’s Chemical Enterprise: Engineered Chemical Products Beyond Commodity Chemicals and Their Distinctive
Features
Commodity chemicals
Compounds/simple mixtures produced in large quantities and for which the price is the differentiating factor.
Example: sulphuric acid.
Engineered chemical products
Type of product Features Example
Specialty chemicals Compounds/simple mixtures produced in small PolyDADMAC used as a flocculant in the
quantity (typically less than 1,000 tonnes/year) drinking water industry.
for which performance is the differentiating factor. Key attribute: molecular structure.
Formulated products Multicomponent (typically 4 to 50 ingredients), Exfoliating gel.
multifunctional (accomplishing more than one Key attribute: microstructure.
function valued by the customer) and/or
micro/nano-structured (with value significantly
deriving from their nano/microstructure) systems.
Bio-based concepts A range of biomaterials, drugs, tissue and metabolic Tasimelteon used to treat the non-24-hour
engineering technologies. sleep–wake disorder in totally blind people.

2
Key attribute: biological activity.
Technology-based consumer goods Consumer goods whose functionality is provided Disposable nappy.
by a chemical/physical technology. Key attribute: materials and assembly.
Devices Devices that carry out a physical/chemical process. Artificial heart valve.
Key attribute: materials and configuration.
Virtual chemical-based products Software supporting chemical process-related Aspen Plus1.
industry activity. Key attribute: computational performance.
CHEMICAL PRODUCT DESIGN AND ENGINEERING 3

Fig. 1. Elucidating the scope of chemical product engineering: a conceptual model struc-
turing the subject (adapted from reference 1).

task of converting costumer needs and/or new technologies into new products. In
practice, this task involves the embodiment of property, process and usage
functions, which are systematised in the chemical product pyramid. This must
be integrated with the design of a manufacturing process. The effective incorpo-
ration of the chemical product pyramid and process design to create a successful
product marketed at a global level demands the adoption of a multifaceted
approach covering several time and length scales. Hence, design of a new chemical
product spans a broad space of knowledge at the interface of chemical engineering
core concepts (such as thermodynamics and fluid dynamics) with other scientific
(for example, biology) and management (for example, marketing) topics.
While stressing a distinction between chemical product engineering and
chemical product design might seem superfluous, such a distinction is important
for chemical product engineering to solidify as an acknowledged branch of
chemical engineering science, education and practice.
Overall, chemical product engineering and design principles aim to bring
systematisation into innovation practice and thus address the specificities of high
added value products for which performance and time-to-market are the success
critical factors (Table 1).
4 CHEMICAL PRODUCT DESIGN AND ENGINEERING

In the remainder of this chapter, an overview of chemical product engineering


fundamentals is provided, organised by the conceptual model presented in
Fig. 1. First, the three main pillars supporting chemical product design are briefly
discussed; then, a template for chemical product design is proposed. Examples are
providedinordertoillustratetheideaspresentedsincegettingintothechemicalproduct
engineering perspective gains a lot by adopting a ‘‘learning by doing’’ approach.

2. Three Pillars Supporting Chemical Product Design

2.1. The Chemical Product Pyramid. The concept of chemical product


pyramid is presented in Fig. 2.
An important trait of engineered chemical products is that customers gener-
ally do not judge their value based on technical criteria, such as the concentration of
a thickening agent, but rather on functionality and performance attributes, like the
ease of application of a lotion. These attributes, used to characterise product quality
from the customer perspective, may be labelled as quality factors.
Quality factors are often inherently qualitative and subjective, and hence it
is useful to translate them into quantitative parameters, the so-called perform-
ance indices.

Fig. 2. The chemical product pyramid: concept and illustrating example (adapted from
reference 1).
CHEMICAL PRODUCT DESIGN AND ENGINEERING 5

Performance indices, and related quality factors, depend on three main


factors: (i) the composition and physico-chemical properties of the materials
forming the product; (ii) the product structure or configuration, which may be
strongly linked to the manufacturing process; and (iii) the product usage
conditions.
The property function conveys the relationship between performance indices
and product composition, materials’ properties and product structure/configura-
tion. As a simplified illustration (Fig. 2), consider a pharmaceutical ointment
containing an active ingredient (11). The capacity of the ointment to provide
controlled release of the active ingredient in order to avoid the need for multiple
applications over the day is a key quality factor. In objective terms, this aspect of
the ointment’s performance can be described based on the concentration of the
active ingredient in the patient’s plasma (performance index). Since the ointment
is an emulsion, the active ingredient’s partition coefficient (dispersed phase/
continuous phase), its diffusion coefficient in the continuous phase and the size
of the dispersed aqueous droplets constitute key variables affecting product
performance. These variables define a mass transfer model that can be used to
describe the transfer of active ingredient to the plasma through a series of
interphase equilibria and diffusion processes, and thus predict its concentration
in the plasma. Such a model works as a property function.
For engineered chemical products, the structure/configuration often has a
dominant influence over functionality and end-use properties. The desired prod-
uct structure/configuration requires the selection of materials, but frequently it is
largely determined by the manufacturing process. Process functions relate the
process conditions to the structure/configuration attributes of the product. For the
pharmaceutical ointment (Fig. 2), the size of the aqueous droplets dispersed in the
emulsion depend on the mixing rate and time in the emulsification process,
amongst other variables; a droplet break-up model constitutes a process function,
predicting the droplet size characteristics based on the mixing conditions.
In addition to product composition, materials’ properties and product struc-
ture/configuration, the circumstances under which the product is used also affect
perceived quality. Parameters describing the way the customer uses the product
(such as the frequency with which the patient actually applies the ointment, the
quantity applied and the skin area covered) and environmental conditions under
which product usage takes place (for example, the blood volume of the patient)
cannot be directly controlled. Under the Taguchi nomenclature, these parameters
correspond to noise factors affecting product performance and hence demanding
robustness from the product (12,13). Usage functions relating performance indices
to customer interaction parameters and usage environmental conditions may be
established. For the pharmaceutical ointment, a dynamic model able to predict the
concentration of the active ingredient in the plasma depending on the values
assumed for specific usage conditions will be an example of such a function (Fig. 2).
The idea of a chemical product pyramid (Fig. 2) has been introduced to
systematise the relationships between product composition, materials’ physico-
chemical properties, process variables, product structure/configuration attrib-
utes, usage variables and product quality (1). The base of the three-sided pyramid
is defined by the materials, process and usage spaces, which determine the
chemical product space occupying the top of the pyramid. Connections between
6 CHEMICAL PRODUCT DESIGN AND ENGINEERING

and within these four spaces are established by property, process and usage
functions.
Overall, the chemical product pyramid embodies the technical core of
chemical product engineering — in practical terms, the subject is concerned
with the development of property, process and usage functions and their applica-
tion to design and manufacture chemical products with end-use properties valued
by the customer.
2.2. Chemical Product Design and Process Design Integra-
tion. Since achieving a desired functionality is the core of chemical product
engineering, the connection between chemical product design and process design
can be underestimated (14).
No product can exist without a manufacturing process and, as already
discussed, the characteristics of engineered chemical products often strongly
depend on manufacturing choices and conditions. Therefore, the effective inte-
gration of product and process design is a key to achieving the desired function-
ality and quality.
The development of processes for the manufacture of engineered chemical
products has some differences when compared to process engineering associated
with commodity chemicals. Engineered chemical products tend to have short
market lifetimes and a value much greater than that of the raw materials. This
means that the emphasis on efficient processing is reduced. The phased approach
involving conceptual, basic and detailed designs followed by procurement and
construction, which has proved so successful for the design of plants for commodi-
ties, is not suitable for the design of manufacturing processes for engineered
chemical products. Instead, a process development strategy privileging speed over
optimisation is appropriate to deal with time-to-market as a key for success.
Integrated product design and process design is required for this purpose.
The advantages of product design and process design integration have been
illustrated in the literature (15). An illustrative example reports the development
of a cosmetic lotion (16). In this application, an optimisation problem was
formulated by incorporating product quality, as assessed by customers, a model
relating the viscosity of the lotion, an oil-in-water emulsion, to its composition,
and a model linking process design and operation with product composition and
structure. Optimal lotion composition and process specifications were thus iden-
tified, with product and process design decisions strongly interacting and made
together. An objective function addressing both product quality and process costs
was employed. The integrated product design and process design solution was
shown to be superior to that obtained when product design and process design
were handled separately, as well as being quicker.
2.3. Multifaceted Multiscale Approach. The quality, functionality,
and ultimately commercial success, of an engineered chemical product marketed
at a megascale are defined at the nano- and micro-scale of its materials and
structure/configuration as well as at the meso- and macro-scale of the manufac-
turing process and distribution logistics. Therefore, developing a new product
requires the adoption of a multifaceted multiscale approach to effectively cover
this range (1,2,17,18). The ultimate aim of chemical product engineering is the
translation of phenomenological laws and models, expressed by property, process
and usage functions, into commercial product technology. This relies on the ability
CHEMICAL PRODUCT DESIGN AND ENGINEERING 7

to synthesise problems over length and time scales spanning many orders of
magnitude. Such an overall approach is necessarily multidisciplinary in nature,
comprising a mixture of fields, from business, social sciences or even fine arts to
basic sciences (for example, chemistry, physics and biology) and chemical engi-
neering technology.

3. Chemical Product Design

3.1. Overview. As already discussed, chemical product design consti-


tutes the operational side of chemical product engineering, the latter being the
whole body of knowledge supporting the development, manufacture and market-
ing of new products. Within chemical product engineering, chemical product
design is defined as the process by which a product is conceived and chosen to
meet specific market demands and business goals. It can be outlined as a logical
procedure and a framework of tools that bring systematisation to innovation.
While the systematisation of the technical side of new product development is
a relatively recent concern in the chemical engineering sphere, it has been better
and earlier established in other engineering fields, such as the mechanical, elec-
trical and biomedical ones (19). By adapting the formulae employed in those fields,
several frameworks for chemical product design have been proposed (6,20–23).
Here, a six step template is suggested (Fig. 3).Once the project’s mission is
defined, the conception of the new product starts by exploring the needs space to
identify the customer requirements the product should meet. These requirements,
initially expressed in the subjective and qualitative language of customers, are
then translated into more quantitative terms, the product performance specifica-
tions, and target values for these specifications are set. Next, alternative product
ideas that may address the customer needs are generated and evaluated in order
to select the most promising ones. Depending on the nature of the project and the
product being developed, the selection step may have to be integrated with a
testing step. In the last step, the final product specifications are established.
Three remarks should be made about this template.
First, although it is described as a sequence of steps, in general chemical
product design is not strictly a sequential process – in practice, various steps tend
to overlap in time and iteration is often necessary.
Second, this is a generic template. Because of the enormous variety of
chemical products which are possible (Table 1), this template should not be
expected to work perfectly in every case. Instead, it acts as a mental checklist
for organising thinking and practice in creating a new product. Furthermore, this
template is particularly oriented to market-pull products, which are designed to

Fig. 3. A generic template for chemical product design.


8 CHEMICAL PRODUCT DESIGN AND ENGINEERING

fill an identified market opening (19). Examples of market pull products are a
deodorising fabric for sports, a needle free injector for medical applications and an
improved dusting cloth. Nevertheless, the template may be adapted to the context
of other innovation approaches (19), including technology-push strategies, where
the initial stimulant for new product development is not a perceived consumer
opportunity, but an advance in technology, a new invention looking for an
application.
As a final remark, the template proposed here refers to chemical product
design seen as a conception stage, one stage within the whole product develop-
ment process. It is preceded by a planning stage, where the project’s mission is
defined, and it is followed by further detailed design and testing, refinement,
production and marketing stages, these strongly depending on the nature of the
product (19).
Details of the chemical product design steps will be given in the body of this
section.
3.2. Identification of the Customer Needs. The first step in designing
a chemical product consists of understanding what and how important are the
customer needs that a successful product should satisfy (Fig. 3).
Before proceeding, it is worth stressing the distinction between customer
needs and product specifications. Customer needs refer to the problems a product
solves and the functions it performs. They are independent of the particular
product concept the team may decide to develop in later stages of the design
process. Product specifications are linked to the product. In particular, at the end
of the design process, product specifications refer to the product concept selected,
being based on technical and economic trade-offs, the characteristics of competing
products as well as customer needs.
The difference between customer needs and product specifications will become
clearer when the definition of target performance specifications and final product
specifications are addressed. For now, a simplified example may help to illustrate
this difference. Parents want nappies to be able to keep the skin of their babies dry
for a full 12-hour sleep. This is the need a successful nappy should meet – no matter
what type of nappy, what mechanism is used to keep urine away from the baby’s
skin, what sort of materials the is nappy made of; all parents care about is that their
babies are not wet. The related product performance specification would be the
volume of liquid that must be retained by the nappy’s superabsorbent polymer layer;
this performance descriptor is linked to a specific nappy concept – a disposable
nappy rather than a cloth one. The amount of polymer that has to be incorporated in
the nappy so that it absorbs the desired volume is a product design specification,
describing how that specific nappy concept can ensure that customers are satisfied.
Exploring the needs space to design a successful product can be achieved as
detailed below, and it is critical to come up with appropriate products later on.
Collect Raw Information About the Customer Needs. When exploring
the needs space, it should be kept in mind that the product is not for the design
team. The needs that will be listed should reflect the requirements of those who
will ultimately use the product and not the team’s own prejudices. This means
that needs identification will begin by finding the customers. The term customers
is used in a loose sense here. It does not necessarily mean those who will buy the
product, rather those who will benefit from it.
CHEMICAL PRODUCT DESIGN AND ENGINEERING 9

Selecting a set of customers to assess their needs can be complicated when


diverse groups of people can be seen as the customer. Whenever possible, lead
users should be identified and included in that set (24,25). These are people who
are experts in a given type of product and those who will benefit the most by its
improvement; they are very useful to the design team because they tend to be
particularly able to articulate their emerging needs and may have already
invented solutions for their problems. In the case of a market-pull product,
lead users are those who very much depend on existing and competing products.
In the case of a technology-push product, lead users are those most responsible for
the technological advance that has stimulated the product development, including
top researchers in the field.
The consensus among marketing organisations is that the best way to get
information on the customer needs is by face-to-face interviews. Fewer than 10
such interviews risk missing significant information, while more than 50 simply
lead to duplication. If organisations are involved, of course individuals within
those organisations must be interviewed. It is worth talking separately to several
of them; it is always surprising the degree of disagreement present within a
supposedly homogeneous organisation.
In addition to individual interviews, focus groups are sometimes used to help
identify needs.
Exploring the context of use of the product either by passive observation or
adopting a ‘‘be a user’’ attitude may also be very useful to gather information on
customer needs. Passive observation involves watching customers using products
that are similar or related to the one being developed or implementing the task for
which the new product is intended. By adopting a ‘‘be a user’’ attitude, the team
collects data by playing the role of an active customer and user of competing or
related products.

Example A: Nonionic surfactants in washing powder


Design problem. A typical washing powder for clothes cleaning made by a major
manufacturer contains 5% of nonionic surfactants, the remainder being anionic
surfactants, mainly linear alkyl sulphonates. Nonionic surfactants are particularly
effective for the removal of greasy stains.
The company would like to increase their loading to 10% to improve washing
performance. The problem is that nonionic surfactants are sticky liquids while
washing powders must be free flowing granules. Simply increasing the nonionic
surfactant loading using current technology results in agglomeration of the
powder into a sticky mess.
How can the nonionic surfactant loading be increased whilst maintaining the
physical properties of the washing powder?

Solution. The design team must first identify the customers, so that they can ask
what is required.
It is tempting to think of the customers just as those who buy and use the
washing powder, but in this case that is not appropriate. While these should
certainly be listened to, they will mostly tell you they want an inexpensive product
that works well. As a result of numerous interviews with users, the design team
has already decided that an increase in nonionic surfactant loading is desirable.
10 CHEMICAL PRODUCT DESIGN AND ENGINEERING

The relevant people to talk to at this stage are within the company, those who
are going to have to implement changes in formulation: the engineers who run the
powder manufacturing plants. The team should aim to talk to 5 to 10 such people,
to give them a good coverage of the issues involved and highlight differences of
opinion.
A typical interview with a plant engineer might go as follows:

What do you do now? We blend the anionic detergent and solid additives in an
extrusion process. This makes the detergent go white and springy, which we
want. We granulate to get the particle size right, and finally spray on
nonionics and perfumes.
What happens if you spray on more nonionic? It is hopeless. You end up
with a big sticky lump of detergent – it is not a powder any more.
What is good about what you do now? It gives a high density product –
better than spray drying the detergent. It is efficient – we get the solids mixing
and physical structuring of the anionic detergent at the same time. It is easy to
change the additives at the end for different products just by spraying
different amounts.
What is wrong with what you do now? Nothing. We make five different
brands this way. Have done for years. They clean people’s clothes OK. I do not
see the problem. Why does marketing want something new anyhow?
Where do you get your materials from? We buy them from a subcontractor –
a specialist manufacturer. They supply the anionic detergent – it is a yellow
sticky paste when we get it – and the nonionic – it is a thick liquid. Solids come
from another company and perfume from a third. The formulation chemists
tell us what to put in and then we have to find it. We always like to have at
least two suppliers of each material.
This interview starts to show the team the needs – how any solution must fit
into the existing process. It also tells them what other people to talk to – the
formulation chemists and the subcontractors who supply the detergents. They will
then get a fuller picture of what is needed. As they finish their interviews, the next
task will be to organise the information they collected into a list of more coherent
and specific needs.

Interpret the Raw Information in Terms of Customer Needs. The infor-


mation gathering stage just described will result in a hotchpotch of conflicting and
incomplete customer statements – often supplemented by video recordings or
photographs – of varying relevance and practicality. The next task is to interpret
and organize these statements, editing them into a cogent list of needs.
Create a Hierarchy of Customer Needs. Typically 50 to 300 customer
needs have been identified and are listed at this point. Such a large number of
customer requirements is difficult to summarise and handle over the subsequent
design steps. It is thus useful to organise the needs into a hierarchy by eliminating
redundant statements and consolidating those that are identical in meaning, and
then grouping the needs based on similarity.
The hierarchy of needs usually consists of a set of primary needs, each broken
down into secondary needs. When the number of customer requirements is large,
CHEMICAL PRODUCT DESIGN AND ENGINEERING 11

a tertiary level may be used. Often this will correspond to a three-layer structure
of strategic, tactical and operational needs.
As the customer needs are being analysed, the team may decide to drop some
of them, either because they appear impractical or are beyond the company’s
expertise. This, however, should be done cautiously to make sure that no key
market opportunities are missed nor critical needs are forgotten.
Establish the Relative Importance of the Customer Needs. Creating a
hierarchy of needs does not provide any information on the relative importance
customers place on the requirements listed. This sort of information is essential
for the team to make trade-offs when designing the new product, and thus the
needs identification step ends with the ranking of the customer requirements.
The relative importance of the needs may be expressed in terms of numerical
weights (based on the design team’s experience or further customer surveys) or
determined by applying Kano’s method (19,26).
Kano’s method results in the classification of each of the needs as one-
dimensional, must-be or attractive. A need will be in the one-dimensional class
when customer satisfaction is proportional to how functional the product is with
respect to that need; the more functional the product is, the more satisfied the
client is. A need will classify as must-be when the customers are more dissatisfied
if the product is less functional with respect to that need, but customers’ satisfac-
tion never rises above neutral no matter how functional the product is; must-be
needs are needs that are expected to be met by the product, so the fact that they
are met does not raise the level of customers’ satisfaction. A need will be labelled as
attractive when the customers are more satisfied when the product is more
functional with respect to that need, but they are not more dissatisfied as the
product is less functional (Fig. 4). Kano’s method involves having a sample of
potential customers responding to a questionnaire constructed based on a specific
template, and the final ranking of the needs reinforces the idea that they are not
equally valued by the customers. For example, improving the performance of a
product in terms of a must-be need that is already at a satisfactory level will be less
valuable than improving the product performance in terms of an one-dimensional
or attractive customer need - the results of Kano’s method will be very useful in
later stages of the design process when the team has to concentrate on trade-offs in
product performance and should aim to outperform in at least one attractive
customer need.

Example B: A fishy business


Design problem. Fish farming is now a major international business.
One company supplies food for salmon and other farmed species to markets
in Scotland, Norway, and Chile. They wish to improve the performance of their
product in order to secure these growing markets. A group of engineering
consultants has been hired to recommend improvements.
The current product consists of cylindrical pellets made by extrusion of a
mixture of fish meal, fish oil and wheat.
An interview with a fish farmer might go as follows:

What do you do now? We use a water cannon to spray the pellets over the fish
pens. We feed once a day.
12 CHEMICAL PRODUCT DESIGN AND ENGINEERING

Fig. 4. Example B: needs that a food product for farmed fish should meet, ranked based on
Kano’s analysis.

What is good about the current product? It is really easy to deliver – the
water cannon is no work. The fish love it. They grow really fast.
What is bad about it? It is oily. You always get an oil film. This seems a waste
and is not good for the fish. I also worry about the oil breeding fungi, which
gets into the gills of the fish. The big fish always get to the top first and eat
most of the food. We would like to give more to the smaller fish.
Where do you buy your food from? We get 70% from your company and 30%
is local waste. The waste is not so good; it is messy and the fish will not eat it
all – they seem to prefer food with oil in. But it is cheap.
Interpret these needs into a form useful for target product performance
specification.

Solution. The customer needs can be easily inferred from this interview. By
applying Kano’s method it is possible to get to a classification like that in Fig. 4,
which conveys the relative importance of the needs from the customers’ perspective.
According to this classification, having good nutritional value is a must-be
need. As the product is fish food, the customers will be extremely dissatisfied if
they are buying something that is not nutritious for their fish; however getting
fish food that feeds the animals is what they expect and hence this function will
not raise the customers’ satisfaction with the product. Having a hygienic appear-
ance is classified as a one-dimensional need – the more hygienic the product looks,
the more satisfied will be the customers; fish food that does not look hygienic will
result in customer dissatisfaction. Finally, being inexpensive appears as an
CHEMICAL PRODUCT DESIGN AND ENGINEERING 13

attractive need. The fish farmers are used to having to pay for fish food, knowing
that cheap local waste is not as good, so paying for the benefits does not
particularly dissatisfy them. However, they will become more satisfied if the
food product they buy works and is inexpensive.
The team is now ready to continue product development.
The way in which needs are extracted from the customer statements,
grouped, organised and ranked will depend on the product being considered. It
will usually be an easy task if the aim is to improve an existing product. The more
innovative the proposed product, the harder it may be to define the needs.
In any case, while the needs identification is the first step in designing the
new product, it may be useful to return to the customers, perhaps a different
group, to further explore a marked list of needs in later stages.

3.3. Establishment of Target Product Performance Specifica-


tions. At this stage, the needs the product should meet have been well docu-
mented, but they are usually expressed in the customers’ language. The aim now
is to convert the requirement statements into product specifications, which will be
revised and completed later on once a promising product concept is identified
(Fig. 3).
Recalling the idea of a chemical product pyramid (Fig. 2) may be useful to
elucidate such a two-phased approach to establishing product specifications.
The customer needs, and related quality factors, provide a clear sense of
the issues of interest to the customers, but, being inherently subjective and
qualitative, they do not provide precise guidance in developing the product. The
next step is thus to establish a set of target performance specifications spelling
out the exact functionality required for the product. These specifications consist
of performance indices (working as metrics for the quality factors) and the
values those indices must take so that the product meets the customers’
expectations.
It is worth noting that, at this point, the design team is still only concen-
trating on a fraction of the product space occupying the top of the chemical product
pyramid. A solution for the design problem has not yet been sought, and so details
of the product’s structure/configuration attributes and materials cannot be given.
Thus, the specifications established at this stage are performance rather than
design specifications.
Meeting these preliminary performance specifications will depend on the
final product concept, and so they are described as target performance
specifications.
Later in the design process, after a promising product idea is identified, these
performance specifications must be revised taking into account particular con-
straints and trade-offs, and the design specifications that provide a physical
description of the desired product will be established.
Establishing the product’s target performance specifications can be struc-
tured in three steps.
Translate the Customer Needs into Performance Indices. Precise and
measurable performance indices able to objectively convey the quality factors
valued by the customers have to be identified.
14 CHEMICAL PRODUCT DESIGN AND ENGINEERING

In doing so, the design team may find useful the following mental checklist to
ensure that the entire space of needs is covered and all the aspects of the science
involved are considered:

(i) write complete chemical reactions for any chemistry involved;


(ii) make mass and energy balances important to the product’s use;
(iii) estimate the rates of any important changes that occur during the product’s
use.

Ideally, each need should be translated into a single performance index. In


practice, this is not always possible, and several metrics may be necessary to capture
one quality factor. Some needs may be difficult to translate into quantifiable metrics.
In such cases, a panel of customers may be used to evaluate the quality factor.
After preparing the list of performance indices, these should be ranked based
on importance. The performance indices’ ranking usually derives from that of the
related needs. In cases where the number of performance indices is small and
information on their relative weight is critical, conjoint analysis may also be
employed (19,27,28).
Collect Competitive Benchmarking Information. The next step is to
define a benchmark, which is an existing or idealised product against which
the functionality of the product being developed can be compared.
Benchmarking information allows the design team to position the new
product in the marketplace, critical for commercial success.
In some cases, when a breakthrough product is being developed, no obvious
benchmark will be available. It may then be necessary to envisage an idealised
benchmark by analogy with similar products or current solutions for the needs the
new product addresses.
Set Ideal (and Marginal) Values for the Performance Indices. Finally,
the available information is synthesised, and ideal (and possibly marginal) values
for the performance indices are set. These values constitute preliminary specifi-
cations, describing the functionality desired for the product. The team will then
generate product ideas and use these preliminary specifications to evaluate them.
The House of Quality, a Quality Function Deployment (QFD) tool (29,30),
may be a useful tool over the process of establishing target product performance
specifications.

Example C: Water purification for a family


Design problem. The WHO estimates that 1.7 billion people do not have access to a
clean water supply (12,000 children per year die from diarrhoeal diseases – the world’s
biggestcauseofchilddeath).Intheabsenceofmajorcivilengineeringprojects,thepointof
demand is the family unit.
A large nongovernmental organisation has decided to design a water purifi-
cation unit suitable for use by individual families in third world countries.
Interviews with governmental and nongovernmental agencies working in
such countries, and most importantly the lead users, usually the mother, reveal
the list of needs a water purification product should meet. Five such needs are
presented in Table 2.
CHEMICAL PRODUCT DESIGN AND ENGINEERING 15

Table 2. Example C: Establishing Target Performance Specifications for a Water


Purification Product
Selected customer Importance Target
needs Performance indices weights values
1 Produces enough Volume of water produced on 1 >100 L/day
water (Must-be a daily basis
need) Volume of water produced in 0.7 >200,000 L
total over the product’s lifetime
2 Supplies safe Protozoa, bacteria and viruses 1 90%
drinking water removal
(Must-be need)
3 Inexpensive Price 0.9 <$15
(Must-be need) Annual operating cost 0.9 <$5
4 Rural focus Use of power supply 0.9 No
(One-dimensional Availability of consumables 0.8 Local
need)
5 Environmentally Environmental sustainability index 0.6 7
sustainable assessed by the team in a scale of
(Attractive need) 1 (less environmentally sustainable)
to 10 (more environmentally
sustainable)

Next, target product performance specifications must be established.

Solution. The needs can be translated into the quantitative metrics listed in
Table 2. The requirement for the amount of water produced (need 1) was specified
both daily and over the product lifetime. The definition of safe drinking water
(need 2) requires thinking about what it is necessary to remove from non-potable
water. Waterborne diseases are the greatest threat – the product should be able to
deal with protozoa, bacteria and viruses. Toxic materials are also sometimes a
problem, but it is probably not useful to focus on these – the variety is very great,
the problem usually local and a better solution is often preventing discharge. So
the design team decided to focus on disease organisms. As for being inexpensive
(need 3), both price and operating cost should be considered. The rural focus (need
4) means that the product must operate in the absence of a power supply. Also any
consumable should be locally available. While the use of quantifiable metrics
should be the aim when setting the target product performance, sometimes
performance indices that take categorical non-numerical values may be adequate,
such as for need 4. Environmental sustainability (need 5) is a rather vague need,
not easy to quantify. It involves, for example, the need to avoid the consumption of
local resources or the production of damaging waste. At this stage, the design team
considered that evaluation of potential products on a scale of 1 (lower rating) to 10
(higher rating) would be suitable as a metric for this need.
Based on the needs ranking, the team rated (scale of 0 to 1) the performance
indices in terms of their relative importance for the success of the new product.
In order to have reference data to position the new product in the market-
place, the team looked for a suitable benchmark. Chlorination is a sound bench-
mark. It is simple, cheap and well established, but has problems in terms of
supply, use and discharge of chemicals. The new product must be more attractive
than local chlorination if it is to succeed.
16 CHEMICAL PRODUCT DESIGN AND ENGINEERING

Finally, the team was able to set target values for each performance index,
supported by the following analysis.
The first question is how much water needs to be purified. A person requires
about 5 L of drinking water per day, but may use up to 50 L for cooking, cleaning
and washing. A household might be of 10 people. So a minimum reasonable
product performance would be producing 100 L of safe water per day. 5 years
would be a reasonable lifetime for the product, so it should be able to produce at
least 200,000 L.
Regarding microbial removal, there are WHO guidelines and it would be
useful to have these in mind, but to see these high standards as an absolute
requirement might be a mistake. After all, an affordable but slightly less effective
device will be more useful than one which no one can buy. So at this stage the
target performance is set as 90% microbial removal. This specification should be
returned to later in the design and may need revision.
Getting the cost right will be critical. What is affordable is clearly
variable, but the design is aimed at some of the poorest people in the world.
Aid agencies may provide some support, but ideally a device affordable by the
users without aid is best. Interviews in Nepal have revealed that a cost above
$15 or an annual running cost of more than $5 will be beyond the reach of
most of the rural population. This requirement is stringent and one the
design team may not be able to meet, and it should not be seen as an absolute
limitation - what is really being said is that the cheaper the better, this limit
being the target.
Having produced a set of target product performance specifications in as
much detail as possible, they should be examined carefully. Being a result of
individuals’ wish lists, they may be entirely impractical, involving, for example,
huge flows, enormous concentrations or massive costs. If this is the case, they
must be revised to be more realistic. This may lead the team to abandon the project
altogether. If the only way of meeting customer requirements is to break a law of
thermodynamics, product development should stop now. This type of critical
examination is sometimes known as a chicken test after a Canadian method
for testing aero engines for their capacity to fly through flocks of geese. The team is
asking themselves if the project is obviously unrealistic, before committing large
resources to it.
At this stage of product design the exploration of the problem space has been
completed. The design team has produced a ranked list of what customer needs
have to be satisfied and put this into quantitative and scientific terms as far as
possible. The team has also checked that the project aims are not wholly
unreasonable, and ideally a benchmark by which to judge the success of product
development has been identified.
Up to this point, it is important to consciously avoid trying to think of
solutions for the customer needs. It has been the time to define what is to be
achieved without prejudice caused by a preconception of what the product will
look like. If an idea of the product’s nature does already exist, it should be kept out
of consideration until the end of the target performance definition stage. Only now
that well-established criteria for the success of any product are available, ideas for
the product itself should start to be developed – only now the team will leave the
problems space to enter into the solutions space.
CHEMICAL PRODUCT DESIGN AND ENGINEERING 17

3.4. Generation of Product Ideas. Some good product ideas are now
necessary. In principle, the design team needs only one idea, the one that will be
manufactured. In practice, finding that one idea truly worth pursuing requires
that a high volume of ideas are generated – as Linus Pauling stated, ‘‘the best way
to have a good idea is to have lots of ideas’’. Estimates vary from business to
business, but the general consensus amongst product developers is that around 50
ideas have to be generated to get a winning product. These ideas will be subse-
quently analysed and evaluated in order to choose the most promising one for
further development (Fig. 3).
To get the 50 or so product ideas, the design team will depend on people more
than publications. The most important people are those in the team. They will
normally assemble for free-ranging discussions that aim at generating possible
answers. How to run such brainstorming discussions is carefully described in the
literature (eg, 31,32) and so is not detailed here. It is only worth mentioning that
these discussions should initially be noncritical, and that all participants should
be treated as equals – promising new product ideas are sometimes abandoned for
reasons as trivial as that the boss’ spouse dislikes them.
Dozens of techniques are available to stimulate creative thinking, either
integrated in brainstorming sessions or applied independently as detailed else-
where (33–35). Natural product screening (36), combinatorial chemistry (37) and
CAMD (Computer Aided Molecular/Mixture Design) approaches (38) can assist
idea generation when the design problem requires finding compounds, formula-
tions or materials possessing a desired functionality.
In addition to depending on themselves, the design team should pay special
attention to customers already using existing, related products. Some of these
customers, the so-called lead users of the business literature, may have already
adapted existing products for their own uses. These lead users often have
excellent suggestions. Other human sources – consultants, faculty, private inven-
tors and the like – may also be approached.
Literature has variable value. Patents and trade information from competi-
tors is often more useful than archival literature.
A wide range of possible sources of product ideas should be explored. Still, the
key is most often the design team.

3.5. Selection of Product Ideas. The team now has 50 or so ideas of


widely varying quality and completeness. These ideas must somehow be eval-
uated to select the most promising for further development (Fig. 3).
Evaluating all ideas in the same detail will require more resources and will
take much longer than the team will normally have. Thus idea selection is better
managed gradually in three stages. First, without quantification, the team should
try to sort the ideas on completely qualitative grounds, reducing the number to
perhaps 20. To large extent, this first selection stage overlaps the end of the
previous idea generation step. Then, with a bare minimum of quantification,
the surviving ideas should be screened to get just the five or so the team thinks are
the best. Later, through much more careful assessment and scoring, the winning
product idea will be selected from these five. Rather than absolute numbers, the
50, 20 and five should be seen as order of magnitude averages.
18 CHEMICAL PRODUCT DESIGN AND ENGINEERING

Ideas Sorting to go From 50 to 20 Ideas. To reduce the number of ideas,


the team starts by making and going through a list of all the ideas.
Redundancy can easily be removed. Often this redundancy will occur because
some ideas are special cases of others. For example, in a discussion of better barriers
for landfills, one idea could be ’’the barrier should capture mercury’’. A second idea
might be ’’the barrier should adsorb all heavy metals except calcium’.
In addition to removing redundancy, ideas that are obvious folly should be
dropped. In doing so, the team should be cautious, because some silly ideas may
contain feasible dreams. Sometimes, it is useful to keep a separate list of these
flawed dreams, just to serve as a stimulus to later development.
Normally, the efforts to remove redundancy and folly will still leave up to 40
ideas. To reduce the number of ideas further, they should be organised into
categories, in a type of outline. How this should be done depends on the ideas
generated – there seems to be no general strategy, ‘‘the material will tell you’’.
Once this outline is made, it may expose gaps, which may imply further brain-
storming. More often, the team will find that large groups of ideas will be
incompatible with the company’s objectives or its strengths. These groups of
ideas can be dropped, a major simplification.
These last steps commonly cut the number of ideas to meet the target of 20.
One note of caution, however, many of the best ideas will often cluster under a
single heading. Not to overspecialize too soon, at least one idea beyond this cluster
should be chosen for the next stage.

Example D: Multilayered polymer films for secure documents


Design problem. Counterfeiting documents is a big business. In the United States
alone, more than $250 million of counterfeit currency is recovered each year. To
reduce this problem, some nations have gone to composite polymer films produced by
multilayer extrusion as an alternative to paper for printing money. Australia has led
the way. Small denomination bills circulate so rapidly that paper bills wear out in
less than oneyear,but large denomination bills circulatemuch more slowly.Because
composite polymer bills last four times longer, their higher cost –twice that ofpaper –
can be recovered in a couple of years.
A given company hopes to build on its experience in multilayer extrusion to
make new films for all types of secure documents. The design team is now devoted
to identifying these documents. What should they recommend?

Solution. The team quickly identified over 200 possible products. After the
redundancy and folly were removed, the ideas could be organised under the
four headings shown in Table 3. Because this list is so broad and often so vague, it
should be further edited to represent the company’s strengths.
Ideas Screening to go From 20 to Five Ideas. The 20 or so surviving
ideas must now be reduced down to a still smaller number, normally five or fewer.
There are still not resources to make more detailed calculations for the 20
survivors, so the design team needs approximate but quantitative tools that let
them continue the screening quickly, but on a more rational basis.
One commonly effective method for this screening is to choose five or fewer
key attributes shared by most of the surviving ideas, and then use them as
CHEMICAL PRODUCT DESIGN AND ENGINEERING 19

Table 3. Example D: Ideas of Products Based on Multilayered Polymer Films


I. Minor improvements in existing products
These improve or modify the existing paper document
A. Currency
1. Paper currency with irremovable polymer strip
2. Machine-readable currency to tell denomination or lifetime
B. Passports and corporate identification
1. Identification page of a passport made of multilayered polymer
2. Fingerprints added with optical ink
C. Credit cards
D. Surface treatments
1. Antidestructive
a. Multilayer polymer window that diffracts light
b. Polymer lamination
2. Antigraffiti coating
3. Miscellaneous
a. Fluorescent patches
b. Antibacterial agents
c. Waterproof coatings
II. Change in substrate
These replace the paper with a polymer composite
A. Synthetic polymers
1. Currency, stock certificates, and Traveler’s checks
a. Synthetic plastic currency stamped with optical ink
b. Polymer currency with different colours for the different denominations using
dyed injected polymers
c. Metal coins that have a polymer core
2. Identification cards
a. Polymer passports
b. Polymer social security card
c. Oriented polymer paper
d. Polymer currency with layers of different orientation
3. Memorabilia like baseball cards made of polymer
B. Optically unique synthetic polymers
1. Currency, stock certificates and traveller’s cheques
a. Polymer films with a multi-layer window
b. Polymer films with different thickness to induce colour
c. Polymer films with partial burnouts to generate patches of colour
d. Machine readable strips for use in vending machines
2. Credit Cards made with a homogenous polymer backing
3. Identification Cards, including driver’s licences
a. Cards
b. Cards that change colours with light
c. Cards with burn out patches that change colour with light
C. Combinations of synthetic and optically unique polymers
III. Adhesive additions
These products use adhesive to attach them to existing documents
A. Currency
B. Documents
1. Multilayer optical films to authenticate legal documents
2. Multilayer films to replace notary stamps
C. Identification
1. Multilayer optical polymer films adhered to passports
2. Multilayer optical polymer films adhered to student IDs with school logo
D. Creditcards
E. Cheques (continued)
20 CHEMICAL PRODUCT DESIGN AND ENGINEERING

Table 3 (Continued)

IV. Environmentally benign products


This smaller set of ideas involves recycling, and is partially redundant
A. Secure products from recycled plastic
B. Secure products with later uses
C. Creditcards

evaluation criteria. These attributes will include factors like scientific maturity,
ease of engineering, risk of failure and cost, as well as key customer requirements
and performance aspects. Factors that are different for different product concepts
should be chosen. For example, even if safety is the most important product
attribute, nothing is gained by choosing safety as an evaluation criterion if all the
potential products are equally safe. All product ideas that are still being consid-
ered should be capable of satisfying all attributes at least to a limited extent. If
there are attributes which are truly essential, all ideas that cannot satisfy this
constraint should be dropped, and the evaluation continued for the survivors in a
so-called concept screening matrix based on the Pugh method (19,39).
A benchmark against which the product ideas will be evaluated has to be
chosen. It can be a competing product, an earlier generation of the product being
developed, any of the product concepts under consideration or a combination of
subsystems assembled to represent the best features of different products.
On the basis of the evaluation criteria chosen, the product ideas will be
compared to the benchmark. All product ideas being considered in the screening
matrix should be described at the same level of generalisation and in similar
language for unbiased comparison. It is convenient to assign the benchmark
ratings of 0, and then assess the alternative product ideas as being better (þ), the
same (0) or worse () than the benchmark in terms of each of the evaluation
criteria. At this stage of the design process, the product concepts tend to be only a
general notion of the ultimate product, and therefore more quantitative scores
would be largely meaningless. In fact, such scores could reduce the focus on
creativity required to develop better concepts in the iterative process of idea
selection and should be left for later. After assessing all the product ideas in the
screening matrix, they are ranked based on the respective net ratings (number of
‘‘þ’’ minus number of ‘‘’’).
The products with the highest ratings in the screening matrix are those that
are chosen for further development. However, before proceeding, it is useful to
explore the possibility of improving the concepts under analysis. This may be
achieved in two ways: by marginally modifying generally good concepts whose
overall quality is penalised by only one bad feature, or by integrating two or more
product ideas to preserve their strengths and make up for their weaknesses.
Possible improved product concepts should be added to the screening matrix and
evaluated along with the original ideas.
The concept screening approach is illustrated by the following example.

Example E: Lab-on-a-chip
Design problem. This example is of a technology-push project.
CHEMICAL PRODUCT DESIGN AND ENGINEERING 21

Developments in the microelectronics industry have resulted in the fabrica-


tion of very small devices becoming standard practice. Channels down to 100 mm
can be cut, with flow control devices and separation stages on a comparable scale,
which means that it is possible to produce miniature chemical plants on a scale of a
few centimetres. While the potential of such devices as detectors or reactors is
large, few commercial products have yet been produced. Prototypes for DNA
sequencing, blood testing and a handful of other applications have been made.
A company holding patents in nanofabrication is looking at the best areas in
which to launch its technology in the marketplace.
The following needs have been established:

Must-be needs
 Use of current technology for a more cost effective product or a completely new
product
 Quick to market (relies on using existing technology)
One-dimensional needs
 Environmentally benign
 Easily marketable
Attractive needs
 Use company’s technology to its full potential

Explore ideas for products that could be produced using the existing technology.

Solution. After brainstorming, the design team came up with 73 ideas; these
were reduced to 39 by the removal of redundancy, folly and excessive require-
ments. The remaining ideas were then organised as shown in Table 4.
This list must be cut to a handful for detailed consideration later. This
reduction can be done in two stages.
The following analysis can be done first. One major identified need is using
existing technology to ensure a swift entry into the market, which is desirable for the
usual reason that the first product into a market typically takes the lion’s share of
sales, even in the absence of other competitive advantages. Also, in this particular
case, the product being considered is intended not just as a product in its own right,
but also as an exemplar of the many other potential applications of the company’s
nanotechnology – a flagship for the company’s future. It is doubly important to
minimize the risk of technological failure or delay. The company’s technology is
currently not developed in the areas of detection and separations. The whole
categories of E, F, G, H and I are therefore rejected not because they are inherently
bad ideas, but because they will not provide the most risk-free and speedy route to a
successful first product. The team may well return to these ideas in future years.
To screen the remaining 21 ideas, the team very crudely assesses each of
them on three criteria: market size, maturity and reliability of technology, and
likely time-to-market. Only those ideas that look promising in all three of these
vital criteria will be considered further. An example screening of the first four
ideas is shown in Table 5. Of these only the first is taken forward to the next
selection phase, as are B1, C1, D1 and J3.
22 CHEMICAL PRODUCT DESIGN AND ENGINEERING

Table 4. Example E: Grouping of Selected Ideas for Lab-on-a-Chip


Detectors/analysers
A. Medical/laboratory
1. Compactanalyticalinstrument
2. Drugdevelopment
3. Screening of blood/vaccines/food/water supplies for infectious agents
4. Diagnosis of disease, eg, acquired immune deficiency syndrome
(AIDS), hepatitis, cancers and Alzheimer’s disease
B. Industrial
1. Chemical plant on-line testing
2. Effluent testing (chemical plant/sewage plant)
C. Commercialenduser
1. Air pollution detector (smoke/CO/H2S)
2. Food testing (nut/GM/rot/animal products)
3. Accuratebreathalyzer
4. Blood sugar testing for athletes/diabetics
5. Homepregnancytests
6. Check blood–iron levels for anaemics
7. Allergytestingatrestaurants
8. Alertingasthma
9. Check pollution levels for surfers
D. Regulatory/policeuse
1. DNA analysis and fingerprinting (spit, urine, hair)
2. Drugdetection – sniffer chip
Producers
E. Medical
1. Hormoneproduction
2. Makinginsulin
3. Waspstingremedy
4. Timed production and dose of drugs, especially for the elderly
5. On-the-spot antidotes for biocides or rabies etc
6. Production of drugs with very short shelf life in, eg, first aid kit
7. Internal surgery – nutrients for tissue growth
F. Industrial
1. Production of specialty chemicals as/when needed
2. Chips to spin fibres – tiny spinnerettes for polymers
3. Manufacture of dangerous chemicals in smallquantities (eg, phosgene/hydrogen cyanide)
G. Commercial
1. Make-up manufacture
2. Night-clubs – smoke/foam production
Combined detectors and producers
H. Commercialenduser
1. Suncream detector andapplicator
2. Variable strength deodorant – alter strength
3. Air freshener – variable strength
4. Worktop coating that tests and releases antibacterial agent
Cleaners/removers
I. Medical
1. Minidialysis – remove alcohol from system to sober up quickly
2. Arterycleaner
J. Industrial
1. Critters-on-a-chip – detect, map anddigest environmental pollutants
2. Cleaning water supplies via ion exchange
3. Fouling/corrosiondetection/remediation
K. Commercial
1. Stain removal for washing – seek out and remove dirt
CHEMICAL PRODUCT DESIGN AND ENGINEERING 23

Table 5. Example E: Screening of Four Ideas for Lab-on-a-Chip as an Illustration


Four product concepts for illustration purposes
A.3. Screening of
A.1.Compact blood/vaccines/food/water
Selection analytical A.2. Drug supplies for A.4. Diagnosis of
criteria instrument development infectious agents disease
Market size þþ þþ þ þþ
Maturity and þ þ þ 
reliability
of technology
Likely time-to- þ   
market
Net ratings þ4 þ2 þ1 0

Ideas Scoring to go From Five to One Idea. With a handful of good ideas
remaining, the team must next choose from these the best one to take forward for
product development.
Because all the remaining ideas are promising, this decision involves con-
siderably more effort than that put into cutting the number of ideas down in the
screening stage just described. As far as possible, the team must quantify how
each idea will measure up to the performance indices that have already been set
for a successful product. This quantification will involve making estimates based
on chemistry and engineering, and perhaps doing some simple experiments. At
the same time, the team wishes to develop the product as quickly as possible and
does not want to put resources into exploring ideas that will end up being rejected.
The key to success in this stage of idea selection is to make reliable choices with
minimal effort.
The exploration of the problems space in the early stages of product design
(steps 1 and 2 in Fig. 3) provided a list of criteria by which to judge the success of
product development. The importance of quantifying the needs, establishing
target product performance specifications, has been emphasised. This approach
will now show its value.
The first step in the final evaluation of the 5 or so remaining ideas is to
estimate how each one will perform relative to the performance indices into which
the customer needs have been translated. In order to do this, the team must gather
more information about each idea. This process will involve firming up exactly
how each idea will work – the team may need to do some simple tests to achieve
this (Fig. 3); they will certainly need to explore the literature some more. As more
detailed information on each idea is generated, the idea itself will change and
crystallise. Thus there is iteration between the idea generation and idea selection
steps – as the team explores an idea in more detail, the idea evolves and new ideas
may emerge. As stressed before, although Fig. 3 presents product design as a
linear six step procedure, this should be understood as a simplification.
Having made an assessment of each idea against the performance metrics,
the team must now make an overall comparison among the ideas. In some cases,
particularly where the customer needs are primarily technical, this will be easy
once good estimates of product performance are available. In other cases, subjec-
tive judgments will be necessary, either in comparing unlike criteria or inherent
24 CHEMICAL PRODUCT DESIGN AND ENGINEERING

in the criteria themselves. In such cases, the design team often proceeds by
drawing up a decision matrix in the same way as was described above for the
screening ideas stage.
Previously, the team was interested in eliminating the weak ideas and the
emphasis was on making quick decisions. Now, they are considering strong ideas
and the emphasis must be on making a good decision. While the methodology of
the decision matrix remains the same, a lot more effort must be put into
evaluating each idea against each criterion, and thus the matrix used now is
usually labeled as a concept scoring, rather than screening, matrix.
The evaluation criteria in the scoring matrix are not only more refined than
those used in the screening matrix, but also have assigned weighting factors
(19,40).
On the basis of the evaluation criteria and their weighting factors, the
product ideas are scored relative to an appropriate benchmark. It is convenient
to assign 5 to the benchmark with respect to all evaluation criteria, and then use a
scale from 1 (poor performance) to 10 (excellent performance), more objective and
finer than that used in the previous screening matrix, to judge the potential of the
new product concepts. An average weighted score is calculated for each product
idea, and the product ideas are ranked.
Before closing the evaluation based on the scoring matrix, it may be useful to
conduct a sensitivity analysis by varying the criteria’s weighting factors and the
concepts’ ratings to ascertain how robust the matrix-based selection is.
As a final note, it should be recognised that the decision matrix is a highly
imperfect tool for making a complex decision, where numerous subjective and
objective criteria are involved. However, this tool has value because it ensures
that all important factors are explicitly examined and because it allows more
detailed research to be conducted where the decision remains inconclusive. Also,
for those technically trained, the decision matrix is a useful way of ensuring that
subjective factors, such as the look of a product, are not overlooked in favor of the
more objective factors that they feel more comfortable with.
Example F: Selecting a polymer film that stops UV light
Design problem. On summer days, cars get hot because UV sunlight passes
through the windows and is absorbed by the car’s interior. Because a given company
has considerable expertise in thin film technology, they are interested in making a
film which could be attached to these windows and block light.
Such a defect-free film should have four key performance properties:

 It should be transparent to visible light


 It should block 99% of UV light
 It should be 70–150 mm thick, but cost less than current competitors
 It should be easy to apply

These properties derive partially from a competitive benchmark, a clear


plastic film with a vapor-deposited metal coating. This benchmark, available in a
variety of colors, costs $8–12 m2.
After idea generation, sorting and screening, a design team came up with two
alternative products. One is a multilayered polymer composite with an internal
CHEMICAL PRODUCT DESIGN AND ENGINEERING 25

UV-absorbing layer and an adhesive backing. The second, more imaginative


product is also a multilayered film, but the internal layer has slight electrical
conductivity – when no current is flowing through this conducting layer, it is
opaque, but when a current is flowing, it becomes transparent.
Which product idea should be developed further?

Solution. After considerable discussions, the design team decides on four attrib-
utes for evaluating these products: cost, engineering, ease of application and
aesthetics. ‘‘Cost’’ should include manufacturing and development expenses.
‘‘Engineering’’ reflects the ease of manufacturing the product. ‘‘Ease of
application’’ includes the effort of installation and maintenance required by the
customer. ‘‘Aesthetics’’ includes both quality and market appeal.
Using these criteria, the design team comes up with the concept scoring matrix
shown in Table 6. The UV absorber with adhesive backing is cheap and easily made
with the company’s technology, but the product is not otherwise much different to
the benchmark. The electrically activated UV absorber is costly, hard to make and
difficult to install. While it is superior aesthetically, it ranks below the benchmark
that it is designed to replace. The company should make the product which builds on
their current skills, if they decide to make any product at all.
Getting Close to a Decision. At this point the design team should have a
good indication of which idea looks the most promising in terms of fulfilling the
customer needs and reaching the target performance defined in the first stages of
product design. However, before proceeding with the development of this idea, it is
important to pause and consider two factors that so far have been largely over-
looked: intellectual property and risk.
Intellectual property is a complex area and should be referred to an expert.
The important point here is to make sure that the ownership of the intellectual
property is clear before large resources are invested. Often the profitability of a
product (notably a pharmaceutical) depends on the exclusive license granted by
patent protection. In such cases the team must ensure there is at least a good
prospect of obtaining such protection before proceeding. In all cases, they must
ensure that the company’s activities will not be restricted by any intellectual
property held by others.

Table 6. Example F: Scoring Two Alternative UV Barrier Films


Product concepts
UV absorber with Electrically
Benchmark adhesive backing activated absorber
Weighted Weighted Weighted
Selection Weighting Rating score Rating score Rating score
criteria factors
Cost 0.25 5 1.25 7 1.75 2 0.50
Engineering 0.25 5 1.25 9 2.25 2 0.50
Ease of 0.25 5 1.25 5 1.25 3 0.75
application
Aesthetics 0.25 5 1.25 5 1.25 9 2.25
Total score 5 6.5 4.0
26 CHEMICAL PRODUCT DESIGN AND ENGINEERING

In assessing how well the product ideas measure up to the criteria set by
customer needs and related target product performance, the issue of risk has been
largely ignored. However, the ideas the team is choosing among may range from
minor developments on an existing product to risky and untested new technology.
This needs to be factored into the final decision.
Risk may take three forms: the product may not work; the product may take
a long time to develop; and external problems may occur because of local politics,
fashion or a changing economic situation. The first of these, product function,
should be unlikely; by this stage the team should have eliminated product ideas
that are likely to fail. The second, development time, is largely a technical issue
that the team should be able to make a good stab at predicting. It can often be
translated into a financial risk – the longer a development program is likely to be,
the greater the uncertainty in cost and the larger the return must be. The third,
external problems, is the hardest to estimate, and ultimately comes down to a
matter of judgment. Nevertheless, at this point, the team should at least think
hard about what factors could compromise the success of the product.
In thinking about risk, one needs to consider both the probability of an event
and the seriousness of its consequences. It is often useful to do this by drawing up a
table giving total risk as a product of probability and consequence. Just as when
using decision matrices, the numbers here should not be treated with excessive
reverence – the procedure is really a means of ensuring all factors are carefully
considered.
Having identified the risks involved in the favored product idea, the team
has three possible responses. They might decide an idea, while attractive, is too
risky to merit time and effort. This realisation might lead them to select a different
idea of smaller potential, or to abandon the project altogether. It is essential to
carefully consider this latter option before proceeding to further development and
manufacturing, when large amounts of money must be invested. Most practition-
ers of product design say that the most common mistake is to abandon projects too
late. Alternatively, the team may decide simply to accept the risks and to proceed.
This minimizes the time to product launch and is often the appropriate strategy if
financial risks can be offset against the advantage of getting to market earlier.
Proceeding in spite of risk is particularly apt where the company holds a patent.
The team should be more circumspect in dealing with risks of a safety or environ-
mental nature. Third, the team may decide to do a little more research, perhaps
including an experimental program (step 6 in Fig. 3), before committing resources
to further product development and manufacture, which is closer to the tradi-
tional approach of prototyping - this will often result in a better product, but also
in a longer (and more expensive) development period. Under these circumstances,
it is convenient to come up with rough rapid prototypes as soon as possible in order
to ‘‘feel the product’’ and get an idea how customers may react to it.

Example G: Moderate scale oxygen production


Design problem. Oxygen, one of the largest commodity chemicals, is often made at
the site where it is needed. The traditional manufacturing method, dating from Von
Linde’s 1905 process, is cryogenic distillation. This technique is extremely effective,
but requires a large capital investment, and so is most suitable for large scale. In the
last few decades, pressure swing absorption has been developed to produce oxygen at
CHEMICAL PRODUCT DESIGN AND ENGINEERING 27

Table 7. Example G: Cost Estimates for Different Column Internals


Column Cost per Column Column Column
internals column volume capacity size cost
Sieve trays 1 1 1 1
Structured packing 2 3 1/30 2/3
Nonselective membranes 10 50 1/50 1/5

smaller scales, capturing this part of the market from distillation. For example,
pressure swing adsorption units not much larger than a beer can are commercially
available to produce oxygen enriched air for patients with emphysema.
A manufacturer of distillation equipment is considering new technologies for
making 85% oxygen at a rate of 6000 scfh. Their current technology, based on high
capacity trays, is not economic at this small scale. As alternatives, they are
interested either in structured packing or in hollow fibre membranes. The
structured packing, which recently came off patent, consists of metal guides,
mounted crossways, looking like stacks of stainless steel venetian blinds. It is an
established technology now supported by both the original manufacturers and
new entrants. Hollow fiber membranes are much more speculative. Selective
membranes are commercially used to produce nitrogen enriched air. Selective
membranes, which retain a nitrogen enriched waste and permeate an oxygen
enriched product, are not commercially attractive at this scale. Porous, non-
selective membranes are not used to affect selectivity but to control a condensate
flow moving countercurrently to the vapor, in a configuration like that of a shell-
and-tube heat exchanger. Such membranes are completely untried, though
academic reports promise productivity increases of several orders of magnitude.
Which, if any, technology merits development?

Solution. The cryogenic distillation of air involves three major types of equip-
ment: compressors, heat exchangers and distillation columns. The compressors
are the biggest capital expense, perhaps one-half of the total; the heat exchangers
are another 30%. Thus one could conclude that distillation design is not important
anyway. However, the compressors and heat exchangers will be standard to any
distillation process, and margins for a commodity like oxygen will always be small.
Therefore cutting the size of the distillation columns could be a significant gain.
To examine the effect of column size in more detail, the equipment capacity
and cost for trays, structured packing and membranes are compared in Table 7. The
first column of figures in this table gives the relative cost of the three internals. In
the past, structured packing has been expensive, but with patents expiring, costs
have dropped significantly. Membrane costs are a complete guess: while in principle
there is no reason that membranes should be more expensive than trays, one
expects that they will need frequent replacement. The other columns in the table
give the capacity, the size and the column cost. The capacity of the membranes is
extremely high, a consequence of the fact that membrane units are largely
unaffected by flooding. The result is that the cost of the membrane-based column
is potentially much less than those of the other internals.
The difficulty with the membrane alternative is that it is so risky. To
understand the origin of this risk, one can compare structured packing with
28 CHEMICAL PRODUCT DESIGN AND ENGINEERING

Table 8. Example G: Risks for Alternative Column Internals


Probability Consequence Risk
Structured packing
Competition from other manufacturers 0.9 0.3 0.27
Performance failures 0.1 0.5 0.05
Nonselective membranes
Competition from other manufacturers 0.1 0.5 0.05
Performance failures 0.7 0.7 0.49

membranes using the risk estimates in Table 8. Structured packing is going to


work. If the company has any trouble, they know that they have not assembled it
correctly, and they can imitate existing products to overcome these shortcomings.
However, the expiration of patents in this area means that many competitors are
now scrambling for new business. Thus the risk for structured packing comes from
the marketplace, not from the technology. As Table 8 shows, the risk for mem-
branes has a different origin. There is no current competition. Even if competitors
appear, they are unlikely to have more experience, so the company will not have
any implicit disadvantages. At the same time, they may have major difficulty in
achieving the performance expected from the literature. If they do have trouble
with membrane performance, they have little experience of how to fix this, so the
consequences will be major.
In this situation, the selection is unclear. The company can do nothing,
accepting that the market for their trays may get smaller in the future. They can
decide that they are efficient manufacturers, and enter the competitive structured
packing market, recognising both the risk and the fact that they cannot do much to
reduce it. Finally, they can gamble on the membrane market, trying through
further technical development to reduce the probability of performance failures
from the current value of 0.7 to a reasonable goal of 0.2. Then – and only then –
membranes may make sense as an alternative to structured packing.

3.6. Establishment of Final Product Specifications. By this stage,


the design team has selected a promising product concept. They are ready to start
refining that concept and manufacturing the product.
But before proceeding, the product performance specifications set earlier
have to be revised – now the team knows the details of the specific product concept
they have chosen and hence can commit to more definite values for the perform-
ance metrics taking into account limitations and trade-offs inherent to that
product concept.
The team has also to describe exactly how the product will be physically, i.e.
to establish the product design specifications (for example, the particle size of the
encapsulated active ingredient in a pharmaceutical cream) that will provide the
desired product performance. The sort of product design specifications that will
have to be defined depend on the nature of the chemical product (Table 1), but
typically the team may need to define chemical compositions, structural charac-
teristics, configuration and physical geometries, materials’ properties, chemical
reactions that occur during the product’s operation and thermodynamics aspects.
The key to a chemical product is often the response to a change, which may involve
dissolution (or precipitation) in a solvent, temperature or other physical variation,
CHEMICAL PRODUCT DESIGN AND ENGINEERING 29

or a chemical reaction (pH change being the most common initiator). It is


important to specify these chemical changes carefully as well as the nature of
the product itself.

Example H: A self-warming baby milk bottle


Design problem. A baby milk manufacturer wishes to market a self-warming
bottle for feeding babies in the absence of an easily accessible power supply. It has
been decided to achieve this by using a double skinned bottle, the inner space being
filled by a material which will undergo an exothermic reaction.
Set relevant product design specifications.

Solution. One must first think what to use for the exothermic reaction. Obvi-
ously, it is much better if it can be regenerated. A quick literature search reveals
the crystallisation of sodium acetate trihydrate from water to be ideal. This
crystallisation can be triggered in a variety of ways, the liquid phase is stable
in the absence of triggering, and can be regenerated by putting the empty bottle in
boiling water. Recent advances in triggering make the reaction much more
reliable and also allow regeneration by microwave heating, ideal for a baby’s
bottle. The maximum temperature reached is 50 C, hot enough to get the milk to
the required 37 C, but without posing any safety hazard.
Two things have to be determined when establishing product design specifi-
cations: the mass of sodium acetate solution required and the interfacial area in
the double skinned bottle.
First, the overall heat balance must be considered:
   
mass of mass of  
þ Cp Tf inal  Tinitial
sodium
 acetate  milk
mass of
¼ DH rxn
sodium acetate

The enthalpy of reaction DHrxn is 125 kJ kg1, and the heat capacities Cp for both
milk and sodium acetate are assumed to be 4.2 kJ kg1 K1. To heat 0.4 kg of milk
from a Tinitial of 15 C to a Tfinal of 37 C, the mass of sodium acetate should be
1.13 kg. This result is quite high, but not impractical.
Next the rate of heat transfer must be considered:

0 1

d UA B
B 1 1 C
C
 ðDTÞ ¼ B þ CDT
dt Cp @ mass of mass of A
sodium acetate milk

where U is the overall heat transfer coefficient from sodium acetate to milk,
around 50 W m2 K1; DT is the difference between the temperature of the sodium
acetate and that of the milk; and A is the surface area between milk and sodium
acetate.
30 CHEMICAL PRODUCT DESIGN AND ENGINEERING

This interfacial area A should be found. To do so the above equation is


integrated, giving:
2 0 13

DT 6 UAt B C7
6 B 1 1 C7
¼ exp6 B þ C7
DT0 4 Cp @ mass of mass of A5
milk sodium acetate

where DT0 is the initial temperature difference between the room temperature
milk, at perhaps 15 C, and the sodium acetate, at 50 C after its rapid crystal-
lisation. The milk should reach 37 C after perhaps five minutes, so 1.5 kg of
sodium acetate should be used, giving a final temperature of 43 C. The milk will be
at 37 C when DT ¼ 7 C, and so:
2  3
0:4
50  A  300  1 þ
7 6 1:5 7
¼ exp6
4  7 () A ¼ 140 cm2
5
35 4:2  10  0:4
3

Thus, the product design specifications are a double skinned bottle, containing
1.5 kg of sodium acetate in the core surrounded by the milk, with an interfacial
area between the sodium acetate and the milk of at least 140 cm2.
The set of revised product performance and product design specifications
established at the last step of the design process (Fig. 3) constitute the final
product specifications, a full description of the product concept to be manufac-
tured. This is the major output of product design as a stage within the company’s
overall new product development process (Fig. 3).
Recalling the idea of a chemical product pyramid (Fig. 2), the team will now
be refining the values of the performance indices; by establishing the product
design specifications, they will also be setting the structure/configuration
attributes, the bottom part of the chemical product space, as well as the
materials space. The usage space should ideally be taken into account in this
process, because the product design specifications should guarantee that the
perceived quality of the product is robust to uncontrollable variations in usage
variables.
Property functions, as defined in the chemical product pyramid (Fig. 2), will
usually be developed to establish some of the final product specifications based on
the required performance (19). The term property function is used in a broad sense
here – it refers to any analytical or physical approximation of the product that
relates performance indices to a set of design decisions. Ideally, at least some of the
performance indices can be analytically modelled. In such cases, three approaches
can be followed to derive property functions (21,41). When the underlying phe-
nomena behind the relationships are well understood, theoretical expressions can
be obtained from a detailed analysis and rigorous modelling. This approach has
been successfully applied to derive functions expressing transport phenomena, for
example. Order-of-magnitude analysis, based on a description of the phenomena
supported by simplifying assumptions, is an alternative approach to obtain
CHEMICAL PRODUCT DESIGN AND ENGINEERING 31

functions when a full scientific elucidation of the system is not available. Com-
parison of causal and opposing effects in solids handling is an example of this
approach. For cases, such as that of systems involving solids, in which the
underlying physical phenomena are poorly understood, empirical models can
be determined through statistical approaches. In some situations, in particular
for complex and elaborated products, no analytical models can be derived at all. In
those cases, it is generally necessary to build physical prototypes in order to
explore the implications of design decisions over product performance (19). Design
of experiments (DOX) techniques may be useful to reduce the number of physical
models that have to be constructed and maximise the amount of information
extracted from each trial (42).
In addition to property functions, usage functions relating performance
indices to customer interaction parameters and usage environmental condi-
tions (Fig. 2) may be useful to establish the final product specifications – they
allow the team to examine the impact of the circumstances under which the
product is used on the perceived quality.
Note that property and usage functions are almost always unique to a
particular product concept. Thus modeling efforts can only be implemented after
the most promising product idea has been selected.
Cost will always be an important aspect to consider when establishing the
final product specifications. So the team may also want to develop and use a so-
called cost model for the product at this stage (19). Cost models are trivial for
simple products, but more elaborated, and also critical, for complex and assembled
goods. They are used to ensure that the product’s manufacturing cost allows the
company and the distribution partners to make adequate profits while still
commercialising the product at a competitive price. The design team will use
the cost model to analyse the impact of design decisions on the cost of the product.
In a sense, it is itself a kind of performance model, but instead of predicting the
value of a technical performance index, it predicts cost performance.
The technical performance models, when available, and at least a prelimi-
nary cost model are used to revise the performance and set the design specifica-
tions for the product. Iteratively, the team comes up with specifications that will
allow the new product to get a good position relative to competitors, satisfy the
customer needs and ensure adequate profits. Competitive maps and conjoint
analysis are tools that may be useful as the team approaches the end of chemical
product design projects (19).

4. Concluding Remarks

During the twentieth century, the chemical industry focused on commodity


chemicals, especially those derived from petroleum. Profits depended on the
efficient, large-scale production of a small number of these chemicals. In recent
decades the emphasis shifted towards the manufacture of higher added value,
engineered chemical products.
The functionalities of these products allow them to be sold at a premium
price, often under patent protection. In contrast, the traditional commodity
products are specified chemically rather than based on performance, and sold
32 CHEMICAL PRODUCT DESIGN AND ENGINEERING

into a highly competitive global market for a wide range of uses. For engineered
chemical products, function is key. For commodities, price is key. These differ-
ences between engineered chemical products and commodities imply important
differences in the way they are conceived, manufactured and marketed. Ensuring
the competitiveness of the profession requires that these differences translate into
an update of chemical engineering practice, education and science.
A chemical engineer involved in the engineered chemical products’ busi-
ness must expect to participate, usually as a member of a multidisciplinary
team, in the analysis of market needs, the creative process of generating product
concepts and the identification of the most promising product idea, in addition to
the manufacturing decisions, which have been the traditional focus of chemical
engineers. Chemical product engineering, as an academic field, addresses these
concerns.
Process engineering dealing with commodities has been the privileged
focus of chemical engineering, being a solidly systematized subject and involv-
ing very effective heuristics to aid teaching and practitioners. Chemical product
engineering demands a perspective different from that of traditional process
engineering. As it increases in importance, it also needs systematization and
heuristics.
In this chapter, a conceptual model for chemical product engineering has
been proposed. This model is presented by analogy to the well-established process
engineering in Table 9. Heuristics for tackling the design of chemical products
have also been outlined here. Chemical products are immensely varied, ranging
from dialysis machines, through herbicides, to ice cream. Inevitably no design

Table 9. Chemical Product Engineering as a Well-Established Field: Analogy to


Traditional Process Engineering
Process Chemical
engineering product engineering
Central Chemical process defined Engineered chemical
concept as a sequence of products defined as
interconnected operations high added value
that transform raw materials goods marketed
into products based on performance
rather than com
position or price
Ultimate Synthesis, design, optimisation, Analysis, development
goal operation and control of and manufacture of
chemical processes products able to meet
customer needs, and
transformation of new
technologies into
commercially viable
products
Operational Process design Chemical product
facet design
Basic Unit operations Property, process and
elements usage functions
Main support Flowsheet Chemical product
for communication pyramid
CHEMICAL PRODUCT DESIGN AND ENGINEERING 33

scheme offers a panacea. Nonetheless a heuristic is now available to provide an


intellectual framework within which one can start thinking.

BIBLIOGRAPHY

CITED PUBLICATIONS

1. R. Costa, G. D. Moggridge and P. M. Saraiva, AIChE J. 52, 1976–1986 (2006),


DOI: 10.1002/aic.10880.
2. J. C. Charpentier, Chem. Eng. Res. Des. 88, 248–254 (2010), DOI: 10.1016/j.cherd
.2009.03.008.
3. K. Wintermantel, Chem. Eng. Sci. 54, 1601–1620 (1999), DOI: 10.1016/S0009-2509(98)
00412-6.
4. R. M. Voncken, A.A. Broekhuis, H. J. Heeres and G. H. Jonker, Chem. Eng. Res. Des.
82, 1411–1424 (2004), DOI: 10.1205/cerd.82.11.1411.52030.
5. P. M. Saraiva and R. Costa, Chem. Eng. Res. Des. 82, 1474–1484 (2004), DOI: 10.1205/
cerd.82.11.1474.52024.
6. E. L. Cussler and G. D. Moggridge, Chemical Product Design (Cambridge Series in
Chemical Engineering), Cambridge University Press, Cambridge, 2011.
7. W. D. Seider and S. Widagdo, Curr. Opin. Chem. Eng. 1, 472–475 (2012), DOI: 10.1016/
j.coche.2012.08.003
8. F. Picchioni and A. A. Broekhuis, Curr. Opin. Chem. Eng. 1, 459–464 (2012), DOI:
10.1016/j.coche.2012.08.002.
9. A. V. P. Gurumoorthy and R. J. B. Smith, Educ. Chem. Eng. 8, e41–e44 (2013), DOI:
10.1016/j.ece.2013.02.004.
10. E. L. Cussler and J. Wei, AIChE J. 49, 1072–1075 (2003), DOI: 10.1002/aic.690490502.
11. F. P. Bernardo, P. M. Saraiva and S. Sim~oes in W. Marquardt and C. Pantelides, eds.,
Computer Aided Process Engineering, Elsevier, 2006, pp. 829–834, DOI: 10.1016/
S1570-7946(06)80148-3.
12. M. Phadke, Quality engineering using robust design, Prentice Hall PTR, New Jersey,
1989.
13. F. Bernardo, E. Pistikopouos and P. M. Saraiva, Comput. Chem. Eng. 25, 27–40 (2001),
DOI: 10.1016/S0098-1354(00)00630-X.
14. T. Stelzer and J. Ulrich, Chem. Eng. Technol. 33, 723–729 (2010), DOI: 10.1002/
ceat.200900478.
15. B. V. Smith and M. G. Ierapepritou, Comput. Chem. Eng. 34, 857–865 (2010), DOI:
10.1016/j.compchemeng.2010.02.039.
16. F. P. Bernardo and P. M. Saraiva in Luis Puigjaner and Antonio Espu~ na, eds.,
Computer Aided Process Engineering, Elsevier, 2005, pp. 1505–1512, DOI: 10.1016/
S1570-7946(05)80093-8.
17. I. E. Grossmann and A. W. Westerberg, AIChE J. 46, 1700–1703 (2000), DOI: 10.1002/
aic.690460902.
18. K. M. Ng in R. Gani and S. B. Jørgensen, eds., Computer Aided Process Engineering,
Elsevier, 2001, pp. 41–54, DOI: 10.1016/S1570-7946(01)80005-5.
19. K. T. Ulrich and S.D. Eppinger, Product Design and development, McGrawHill,
London, 2008.
20. A. W. Westerberg and E. Subrahmanian, Comput. Chem. Eng. 24, 956–966 (2004), DOI:
10.1016/S0098-1354(00)00400-2.
21. C. Wibowo and K. M. Ng, AIChE J. 48, 1212–1230 (2002), DOI: 10.1002/aic.690480609.
34 CHEMICAL PRODUCT DESIGN AND ENGINEERING

22. M. Hill, AIChE J. 50, 1656–1661 (2004), DOI: 10.1002/aic.10293.


23. W. D. Seider, S. Widagdo, J. D. Seader, D. R. Lewin, Comput. Chem. Eng. 33, 930–935
(2009), DOI: 10.1016/j.compchemeng.2008.10.019.
24. E. von Hippel, The sources of innovation, Oxford University Press, New York, 1988.
25. G. L. Urban and E.von Hippel, Manage. Sci. 34, 569–582 (1988), DOI: 10.1287/
mnsc.34.5.569.
26. C. Berger, R. Blauth, D. Boger, C. Bolster, G. Burchill, W. DuMouchel, F. Pouliot,
R. Richter, A. Rubinoff, D. Shen, M. Timko and D. Walden, Cent. Qual. Manage. J. 2,
3–36 (1993).
27. D. A. Aaker, V. Kumar and G. S. Day, Marketing research, John Wiley & Sons, Inc.,
New York, 2006.
28. M. E. Pullman, W. L. Moore and D. G. Wardell, J. Prod. Innovation Manage. 19, 354–
364 (2002), DOI: 10.1111/1540-5885.1950354.
29. Y. Akao, Quality Function Deployment - integrating customer requirements into prod-
uct design, Productivity Press, Cambridge, 1990.
30. J. B. ReVelle, J. W. Moran and C. A. Cox, The QFD Handbook, John Wiley & Sons, Inc.,
New York, 1998.
31. A. Osborn, Your Creative Power, Myers Press, London, 2007.
32. D. Koberg and J. Bagnall, The Universal Traveler, Crisp Learning, Menlo Park,
2003.
33. R. Von Oech, A whack on the side of the head: how you can be more creative, Warner
Books, New York, 1998.
34. A. VanGundy, Idea power: techniques and resources to unleash the creativity in your
organization, American Management Association, New York, 1992.
35. G. Altshuller, And suddenly the inventor appeared-TRIZ, the theory of inventive
problem solving, Technical Innovation Center, Inc., Worcester, 1996.
36. P. Cox and M. Balick, Sci. Am. 270, 82–87 (1994).
37. G. Lowe, Chem. Soc. Rev. 24, 309–317 (1995), DOI: 10. 1039/CS9952400309.
38. L. Achenie, R. Gani and V. Venkatasubramanian, Computer aided molecular design:
theory and practice, Elsevier, Amsterdam, 2002.
39. S. Pugh, Total design: integrated methods for successful product engineering, Addison-
Wesley Pub (SD), 1991.
40. G. L. Urban and J. R. Hauser, Design and marketing of new products, Prentice Hall,
Englewood Cliffs, NJ, 1993.
41. D. Livingstone, Data analysis for chemists-applications to QSAR and chemical product
design, Oxford University Press, Oxford, 2002.
42. D. C. Montgomery, Design and analysis of experiments, John Wiley & Sons, Inc.,
New York, 2004.

FURTHER READING

L. T. M. Blessing, A Process-Based Approach to Computer-Supported Engineering Design,


Ph.D. Thesis, University of Twente, Netherlands, 1994.
U. Br€ockel, W. Meier and G. Wagner, Product design and engineering: best practices, Wiley
VCH, Weinheim, 2007.
R. G. Cooper, Winning at New Products, Accelerating the Process from Idea to Launch,
Addison-Wesley Publishing Company, New York, 1993.
H. S. Fogler, S. E. LeBlanc and B. Rizzo, Strategies for Creative Problem Solving, Prentice
Hall, New Jersey, 2013.
H. Grabowski, S. Rude and G. Grein, Universal Design Theory, Shaker Verlag, Aachen,
1996.
CHEMICAL PRODUCT DESIGN AND ENGINEERING 35

T. E. Graedel and B. R. Allenby, Design for Environment, Prentice Hall, Inc, 1996.
R. M. Kanter, J. Kao and F. Wiersema, Innovation: Breakthrough Thinking at 3M, DuPont,
GE, Pfizer, and Rubbermaid, HarperCollins, New York, 1997.
J. F. Louvar and B. D. Louvar, Health and Environmental Risk Analysis, Prentice-Hall,
New York, 1997.
M. E. McGrath, Setting the Pace in Product Development, A Guide to Product and Cycle-
Time Excellence, Revised edition, Butterworth-Heinemann, London, 1996.
J. McMillan, Games, Strategies, and Managers, How Managers Can Use Game Theory to
Make Better Business Decisions, Oxford University Press, Oxford, 1992.
K. M. Ng, R. Gani, and K. Dam-Johansen, Chemical product design: towards a perspective
through case studies, Elsevier Science, Amsterdam, 2006.
R. M. D. Rosenau Jr. A. Griffin, G. A. Catellion and N. F. Anschuetz, The PDMA Handbook
of New Product Development, John Wiley & Sons, Inc., New York, 1996.
S. S. Stevens, Psychophysics: Introduction to its Perceptual, Neural and Social Prospects,
Transaction Books, New Brunswick, 1985.
R. Turton, R. C. Bailie, W. B. Whiting, J. A. Shaeiwitz and D. Bhattacharyya, Analysis,
synthesis,and design of chemical processes, Prentice Hall, New Jersey, 2012.
J. Wei, Product engineering: molecular structure and properties, Oxford University Press,
New York, 2006.
J. A. Wesselingh, S. Kill, and M. E. Vigild, Design and development of biological, chemical,
food and pharmaceutical products, John Wiley & Sons Ltd, Chichester, 2007.
P. R. Whitfield, Creativity in Industry, Penguin Books Ltd., Harmondsworth, 1975.

R. COSTA
R. G. GABRIEL
P. M. SARAIVA
University of Coimbra
E. CUSSLER
University of Minnesota

G. D. MOGGRIDGE
University of Cambridge

Vous aimerez peut-être aussi