Sei sulla pagina 1di 10

Environmental Management

LIFE CYCLE ASSESSMENT:

A Systems Approach to
Environmental Management
and Sustainability
Use this holistic technique to
identify and quantify the potential impacts
of a product or process throughout its life cycle.

Mary Ann Curran


BAMAC, Ltd.

ociety is becoming increasingly aware that human


activity can have far-reaching impacts. As manufacturing operations have become increasingly diverse,
both technically and geographically from the sourcing
of resources, through manufacturing and assembly, to product usage, and finally to disposal global environmental
awareness has become a business imperative. Companies
and governments alike have started to look at products and
services from cradle to grave.
Examples of a solution to one problem creating another
problem are common (1). Compact fluorescent bulbs
reduce electricity consumption by 75% but come with a
dash of mercury. Biobased fuels reduce greenhouse gas
emissions but contribute to air, water, and soil quality
impacts in the agricultural stage. Reusable cloth diapers,
while made of a renewable, natural material (cotton),
require hot water, energy, and detergents for washing.

1970s
End-of-Pipe Treatment

1980s

Twentieth-century environmental strategies seldom


considered the entire system, and their solutions were not
optimal. Since then, environmental management strategies
have evolved and become increasingly broad in scope,
moving from end-of-pipe treatment to pollution prevention
to sustainable development (Figure 1).
Simultaneously, life cycle assessment (LCA) has become
widely recognized as an effective tool for assessing the
resource use, environmental burdens, and human health
impacts connected with the complete life cycle of products,
processes, and activities. This systems approach enables
decision-makers to identify environmental hot spots, as well
as improve industrial systems without shifting burdens elsewhere. What started as an approach to compare the environmental goodness (greenness) of products has developed into
a standardized method for providing a sound scientific basis
for environmental sustainability in industry and government.

1990s

Waste
Pollution Prevention
Minimization/Reduction
Cleaner Production

2000s
ISO Certification

Sustainable Development

Life Cycle Assessment

Life Cycle
Sustainability Assessment

p Figure 1. Environmental management strategies have evolved from end-of-pipe treatment to sustainable development.

26

www.aiche.org/cep October 2015 CEP

Copyright 2015 American Institute of Chemical Engineers (AIChE)

The product life cycle


Figure 2 depicts the main stages of a products life
cycle; energy, transportation, and waste management are
relevant throughout the life cycle.
Raw material acquisition. This stage includes the
removal of raw materials and energy sources from the
earth, such as the harvesting of trees or the extraction of
crude oil. Land disturbance, as well as transport of raw
materials from the point of acquisition to the point of processing, are considered part of this stage.
Manufacturing. The manufacturing stage produces the
product from the raw materials and delivers it to consumers. Three substages are involved in this transformation:
material manufacture. This substage converts raw
materials into a form that can be used to fabricate a
finished product. For example, several manufacturing
activities are required to produce a polyethylene resin from
crude oil: The crude oil is refined, ethylene is produced in
an olefins unit, and the ethylene is polymerized to produce polyethylene. Transportation between manufacturing
activities and to the point of product fabrication should also
be accounted for, either as part of materials manufacture or
separately.
manufacturing. In this substage, the manufactured
materials are processed to create a product and make it
ready to be filled or packaged. Examples of such activities
include blow-molding a bottle, forming an aluminum can,
or producing a cloth diaper.
filling, packaging, and distribution. This stage
includes all manufacturing processes and transportation
required to fill, package, and distribute a finished product.
Energy and environmental wastes caused by transporting the product to retail outlets or to the consumer are
accounted for in this step of a products life cycle.
Use. Consumers are most familiar with this stage
the actual use, reuse, and maintenance of the product.
Energy requirements and environmental wastes associated
with product storage and consumption are included in
this stage.
Recycling and waste management. Energy requirements
and environmental wastes associated with product disposal
are included in this stage, as well as post-consumer wastemanagement options, such as recycling, composting, and
incineration.
LCA according to the ISO standard
An LCA entails inventorying all the material inputs
from the Earth and the outputs to the environment. The
results of the life cycle inventory are then run through
impact models to calculate potential environmental impact
scores, called impact indicators. The LCA model aims
to cover all activities related to a product or function, all
Copyright 2015 American Institute of Chemical Engineers (AIChE)

effects anywhere in the world, and all relevant substances


and environmental themes, as well as a long-time horizon.
While simple in concept, the conduct of an LCA can
be complicated, mainly due to the large amount of data
needed. Fortunately, the increasing availability of LCA
databases and software programs makes it easier to conduct
an LCA. In addition, the International Organization for
Standardization (ISO) established a protocol for performing an LCA study (2, 3). The ISO 14040 series of standards
provides the framework for an LCA, which consists of four
interrelated phases:
1. Clearly defining the goal and scope of the study
(including selecting a functional unit).
2. Compiling a life cycle inventory (LCI) an inventory of relevant energy and material inputs and environmental releases.
3. Evaluating the potential environmental impacts associated with the identified inputs and releases.
4. Interpreting the results to enable more-informed
decision-making.

1. Define the goal and scope of the LCA


An LCA begins with a clearly stated goal. The goal
helps to establish the study boundaries and guides the data
collection efforts.
The private sector is incorporating LCA in many appliRaw Material
Acquisition
Manufacturing

Disposal
Use

p Figure 2. Life cycle assessment evaluates the cradle-to-grave impacts


of a products life cycle, from the gathering of raw materials from the earth,
through manufacturing and use, to the eventual return of materials to the
earth. The arrows represent transportation.

CEP October 2015 www.aiche.org/cep

27

Environmental Management

cations, including various aspects of product design and


development, manufacturing, marketing, use and reuse,
and disposal and end-of-life management. Common goals
for an LCA include:
establish a baseline of overall environmental impact
to identify environmental hot spots
identify possible opportunities for improvement
across the product life cycle
compare alternative manufacturing processes or
supply chains to identify potential trade-offs.
determine the environmental preferability among
alternative product choices
enable continuous product improvement (often with
a concrete target, e.g., the new product must be x% less
impactful than its predecessor while providing comparable
performance).
Most government agencies and other public sector entities lag behind the private sector in embracing LCA as a
tool to support decision-making. LCA results can be useful
in setting public policy at multiple levels, and can establish a culture based on life cycle thinking that helps set the
course toward a greener, more environmentally sustainable
economy. Common goals for a public sector LCA include:
inform government policies and the prioritizing of
their programs and activities
establish consistent policies across consumers, producers, suppliers, retailers, and waste managers
establish consistent policies and policy goals, such as
harmonizing regulations, voluntary agreements, taxes, and
subsidies
introduce policies that appropriately support postconsumer recycling systems.
The goal setting and scope definition phase also
requires the selection of the functional unit, a unique
feature of LCA that sets it apart from other environmental
assessment approaches. The functional unit is defined by
the service that the system being studied provides and is
further shaped by the studys goal.
Limestone
Mining

For example, a study of compact fluorescent lightbulbs (CFLs) might define the system functionality as the
production, use, and disposal of a single bulb, or it could
refer to a fixed quantity, such as 1,000 bulbs. Although the
results of a study defined in this way may be useful for
identifying environmental hot spots, a different function
is needed to compare CFLs to a competing product, such
as incandescent bulbs. A more-appropriate function for
that type of study would be performance-based, e.g., the
amount of light needed to illuminate a 15-m2 room with
1,000 lumen for one hour.
It is important to properly set the scale of the functional
unit. If it is set too small, the LCA might attribute to it an
inappropriately small (often nearly infinitesimal) share of the
total input to or impact of the system (4). For example, the
amount of mercury used to make a single CFL is relatively
small, so the impact associated with making one, or even a
thousand, CFLs may not be significant. On the other hand,
an LCA with a functional unit of, say, ten million light bulbs
would identify a much larger impact.

2. Compile a life cycle inventory


The life cycle inventory is the compilation of data
about the various unit processes within the system under
study. Typically, a flow diagram, such as the one for a
concrete product shown in Figure 3 (5), depicts the processes that make up the system. Flow diagrams are, in fact,
huge webs of interconnected unit processes that fulfill the
system function and support the functional unit, as defined
by the study goal. In todays era of digital databases, an
LCA study can easily be comprised of several hundred unit
processes.
The ISO standard provides a general framework for
conducting an LCA, but it is open to much interpretation by the practitioner. LCAs can produce different
results even if the same product seems to be the focus
of the study. Numerous factors might account for such
differences, including different goal statements, differ-

Clay
Extraction

Grinding
Aggregates
Kiln and
Calcinator

Clinker

Gypsum

Other Raw
Materials

Grinding

Admixtures
Recycling

Water

Cement

Concrete

Use

Demolition

Landfill

 Figure
p Figure 3. This flow diagram represents the life cycle of a concrete product. Source: Adapted from
(5). 3. This flow diagram represents the life cycle of
a concrete product. Source: Adapted from (5).

28

www.aiche.org/cep October 2015 CEP

Copyright 2015 American Institute of Chemical Engineers (AIChE)

ent functional units, different boundaries, and different


Inputs
Inputs
Inputs
assumptions used to model the data (e.g., the use of cut-off
rules, co-product allocation, assigning credit for avoided
burden, and applying consequential LCA). The key is
Process 1
Process 2
Process 3
to keep assumptions to a minimum and explicitly report
the assumptions and values on which the LCA is based.
Readers of the study can then recognize the judgments and
decide whether to accept, qualify, or reject them and the
Product B

study as a whole.
Product A
Product C
Product D
Cut-off rules. Ideally, a life cycle study would account

for all life cycle steps and 100% of the content of product,
modeled with data for the actual materials and processes.
p Figure 4. Expanding system boundaries is one way to avoid the need to
However, we have an intuitive idea that some processes
allocate inputs and outputs among various products.
will have little impact on the study; efforts to collect data
processes lend themselves to physical allocation, because
about them may be unnecessary.
Cut-off rules are typically expressed in terms of mass.
they involve materials with physical parameters that proFor example: The study will account for at least 95%
vide a good representation of the environmental burdens
of the total mass of inputs, and no input that individually
associated with each co-product, such as mass, reaccontributes 1% or more of the mass shall be excluded.
tion enthalpy, and energy content. As a last resort, when
All cut-off assumptions should be verified through
physical relationships do not apply, the standard allows
sensitivity analysis.
allocation based on other relationships, such as economic
Co-product allocation. Industrial processes are typically
considerations.
multifunctional and produce more than one useful output,
Credit for avoided burden. In the system expansion
making it necessary to partition the process input and
approach, the boundaries are extended to include the alternaoutput flows among the useful outputs. Because this can be
tive production of exported functions (i.e., co-products). If
done in many ways, co-product allocation can be controanother product could be displaced in the market by a coversial. The ISO standards give preference to avoiding
product, the impacts associated with that other product are
allocation, which can be done by modeling subprocesses or
subtracted from the impacts of the subject of the LCA.
by expanding the system boundaries.
For example, a typical corn mill produces corn grain
The first option subdivides a process into subprocesses
together with co-products, such as corn oil. The facilwith inputs and outputs that can be assigned to individual
ity represented by Table 1 (6) consumes 77,228 Btu/gal
co-products. This approach is useful when operating data
of ethanol. Because the corn oil can be marketed as a
are provided for a manufacturing facility as a whole but
replacement for soybean oil, the LCA can take credit for
individual co-products can be traced to separate sub
the energy involved in producing the displaced soybean
processes within the facility.
oil (39,333 Btu/gal), making the net energy impact of corn
The second option avoids allocation by expanding the
grain production 37,895 Btu/gal of ethanol.
system boundaries. The system shown in Figure 4 consists
While the system expansion concept seems reasonable
of three processes (1, 2, and 3) that make four products (A,
on the surface, it can be problematic. It requires the identifiB, C, and D). An LCA will compare A and C as alternacation of a way to produce an alternative byproduct, which
tive ways to achieve the function of interest (lets call that
in turn requires additional market data and other informaalpha). The Process 1 inputs go into both A
and B, but the input data are not broken down
Table 1. The system expansion approach credits corn mills with the amount
by product. By coincidence, B and D provide
of energy that would have been required to make a competing product
(such as soybean oil). Source: Adapted from (6).
the same functionality (beta). Expanding the
system boundary for C allows the LCA to
Energy Use, Btu/gal EtOH
compare Process 1 with Process 2+3 based on
Without
For
With
the same total functionality, a + b.
Ethanol, Co-Products, Co-Product
Displaced
Co-Product
wt.%
wt.%
Credit
Co-Product
Credit
However, allocation might be unavoidable,
such as when the data for the subprocesses or
Output*
48
52
77,228
39,333
37,895
for the expanded system cannot be acquired
* Weighted average of dry mill and wet mill output.
easily. The ISO standard specifies the use of
1,000 Btu/U.S. gallon = 0.279 megajoule per liter (MJ/L).
physical relationships in such cases. Some
Copyright 2015 American Institute of Chemical Engineers (AIChE)

CEP October 2015 www.aiche.org/cep

29

Environmental Management

tion. In addition, if the credit for avoided burdens offset


elsewhere is large, the net impact could be calculated to be
negative (7).
The LCA literature gives special attention to recycling,
especially open-loop recycling. When material used in
one system is recovered, reprocessed, and used in another
application, the second application requires less virgin
material. There are several different ways to assign the
savings and disposal burdens to the systems producing
and using the recovered material. Material production,
collection, reprocessing, and disposal burdens can be
allocated over all the useful lives of the material, using
an approach similar to co-product allocation, based on
the mass, energy, economic value, etc. of the recycled
material. Often, boundaries are drawn between successive
useful lives of the material to share the burdens between
the system that produced the material and the one using
the recycled material. Sharing burdens equally is called
the 50:50 method. Alternatively, production burdens can
be assigned entirely to the first system using the virgin
material; this is referred to as a cut-off approach, because
environmental burdens that occurred before the point of
collection are disregarded.
Consequential LCA. An LCA with expanded study
boundaries that encompass the likely consequences of a
decision is known as a consequential LCA. In the corn

ethanol example, the decision is whether and by how


much to increase the production of fuel ethanol from
corn. Searchinger, et al. (8) used the Greenhouse Gases,
Regulated Emissions, and Energy Use in Transportation (GREET) model developed by Argonne National
Laboratory, which models the release of greenhouse
gases (primarily CO2, CH4, N2O) and reports them as
CO2-equivalents (CO2e), to evaluate this decision. In an
attributional LCA, which does not take into account the
consequences of increased corn production, they found
that the greenhouse gas (GHG) emissions associated with
corn-based ethanol were 20% lower than those of conventional gasoline. Their consequential LCA took into
account that additional demand for corn starch for ethanol
production would lead to higher prices for corn, soybeans,
and other grains, which would induce land-use changes.
Considering these factors, they predicted that the GHG
emissions for corn-based ethanol would be 47% higher
than those for gasoline.
Inventory data availability and transparency. The lack
of readily available inventory data continues to be a major
hurdle for LCA practice. In many instances, creating an LCI
begins with the collection of raw data extracted from various
sources, such as a plants accounting records, national statistics, technical journals, etc., including perhaps some that
might seem unrelated to the process for which the dataset is
being developed. It is typically necessary to call on several

Emissions,
e.g., chlorofluorocarbons
Midpoint

Chemical Reaction
Releases Cl and Br+

Area of
Protection

Climate Change
Ozone Depletion
Human Toxicity

Cl and Br+ Destroy Ozone Layer


Midpoint Measures
Ozone Depletion Potential (ODP)

Human
Health

Respiratory Organics
Ionizing Radiation
LCI Results

Noise
Accidents

Thinner Ozone Layer Allows


More UVB Penetration,
which Leads to Endpoints

Photochemical Ozone
Formation

Natural
Environment

Acidification

Skin Cancer

Cataracts

Crop Damage

Marine Life Damage

Immune System
Suppression

Damage to Materials
(e.g., Plastics)

p Figure 5. Ozone depletion midpoint vs. endpoint modeling in life cycle


assessment. Source: Adapted from (11).

30

www.aiche.org/cep October 2015 CEP

Eutrophication
Ecotoxicity
Land Use
Resource Depletion
Dessication Salination

Natural
Resources

p Figure 6. Impacts can be modeled at the midpoint or endpoint (area of


protection) of a cause-effect chain. Source: Adapted from (10).

Copyright 2015 American Institute of Chemical Engineers (AIChE)

sources to collect a sufficient amount of data.


Commercial LCA software tools make data collection
easier. Some tools, however, do not provide information
about how the data contained within them were modeled,
so they function somewhat as black boxes.
Attempts to create publicly available LCA databases
can be found scattered across the globe, although some
have been more successful than others in terms of the
amount of data and quality of the datasets (7). The Swissbased ecoinvent database has been instrumental in increasing the number of LCA case studies. Full use of the database requires purchase of a license. In the U.S., a national
inventory database is freely available from the U.S. Dept.
of Agricultures LCA Commons website (www.lcacommons.gov). This site also hosts the Life Cycle Inventory
Database created by the U.S. Dept. of Energys National
Renewable Energy Laboratory (NREL).

3. Evaluate the life cycle impacts


Life cycle impact assessment (LCIA) converts the large
quantity of input and output data in the life cycle inventory
into relevant impact category indicators. LCIA modeling is
unlike risk assessment, which integrates data and information across a broad range of activities and disciplines,
including source characterization, fate and transport modeling, exposure analysis, and dose-response assessment.
LCIA does not necessarily attempt to quantify site-specific
impacts associated with a product, process, or activity,
because the inventories are often aggregated in a form
that does not include information about the geographical
location of emissions (e.g., where a product is used). Nor
does LCIA try to capture actual environmental impacts that
have happened or will happen during a products life cycle,
since much of the data are calculated or modeled rather
than directly collected or observed.
In LCIA, the individual impact models convert inventory data into common units, such as CO2e, and aggregate
the converted results within the same impact category, such
as global warming potential, to create an impact indicator score for each impact category. The modeling can
consider impacts either at a midpoint within the causeeffect chain or at an endpoint (911). The more common
midpoint approach involves fewer debatable assumptions
and accommodates less-well-established facts; the endpoint approach provides more-intuitive metrics (e.g., loss
of crops instead of kg CO2e). Figure 5 (11) compares
midpoint and endpoint modeling for the ozone-depletion
cause-effect chain.
The first step of the LCIA selecting the impact
categories to be considered is conducted during the
initial goal and scope definition phase to guide the LCI
data collection process. Figure 6 (10) shows the most

Table 2. Calculating global warming potential (GWP),


an impact indicator.
Species

Quantity, kg

GWP, kg CO2e/kg

CO2

100

CH4

10

25

SO2

common midpoints and endpoints.


For consistency, impact categories are often selected
based on an impact assessment guidebook or its implementation in software. Thus, in practice, many LCA studies
report impacts according to the selected model.
The most frequently used LCIA models include:
EPS 2000; IMPACT World+; LIME: Life Cycle Impact
Assessment Method for Endpoint Modeling; ReCiPe; and
TRACI: The Tool for the Reduction and Assessment of
Chemical and Other Environmental Impacts. ISO does not
specify any particular model.
There are no required impact categories that must be
considered. A task group established under the United
Nations Environment Programme / Society for Environmental Toxicology and Chemistry (UNEP/SETAC) Life
Cycle Initiative (see the sidebars on pp. 3435) is developing guidance to support the coherent development and
application of impact indicators addressing any subject of
stakeholder concern (12).
Energy use is sometimes thought of as an impact category. However, in LCA, the focus is on the impacts that
are the consequence of producing and using energy (e.g.,
electricity) due to emissions (CO2, SO2, etc.) and resource
use (coal, gas, oil). The same reasoning applies to waste.
The focus is on impacts that are the consequence of
waste-management processes (e.g., landfilling, incineration, etc.), not the inlet waste stream to be processed
by those techniques. Some of these categories have not
yet been calculated due to a lack of (proper) data and/or
methods.
The second step in LCIA is to convert the emission data
into the chosen impact indicator. For example, consider
the inventory in Table 2, which contains emission data
for carbon dioxide, methane, and sulfur dioxide. Carbon
dioxide and methane are classified as contributors to global
warming, whereas SO2 is not. Characterization factors
(CFs) are then applied to the greenhouse gases to model
global warming potential (GWP). In this case, the CFs are
the GWP of each individual component. The GWP of CO2
(by definition) is 1 kg CO2e, and the GWP of CH4 is 25 kg
CO2e/kg CH4). The impact indicator for the category of
global warming potential is:
(1 100) + (25 10) = 350 kg CO2e
Article continues on next page

Copyright 2015 American Institute of Chemical Engineers (AIChE)

CEP October 2015 www.aiche.org/cep

31

Environmental Management

4. Interpreting the results


Understanding the results of an LCA can be a challenge because of the vast amount of data, diversity of
physical units, uncertainty in the models, and the need to
make value judgments. LCA reports typically run about
100 pages and are accompanied by numerous spreadsheets
that contain the LCI. As mentioned earlier, it is imperative
that all modeling assumptions and data sources be clearly
reported with the results. Furthermore, it is important to
remember that not all environmentally relevant information
can be quantified; impact data are not available to model
certain inventory data, such as for nanotechnology. If LCIA
models cannot model inventory data, there is a risk that
important input or output data that should be reported in
the final analysis could be omitted. The following paragraphs discuss other considerations for interpreting LCA
results.
No clear winner. The results of an LCA seldom identify
a clear winner among alternatives. In some cases, it may
not be possible to state that one alternative is better than
the others because of the uncertainty in the final results.
This does not imply that efforts have been wasted or
that LCA is not a viable tool for decision-makers. The
Table 3. Example classification and characterization steps
for global warming potential (GWP), human health toxicity
potential (HTP), and ecological toxicity potential (ETP)
of airborne emissions.
Characterization Factor
GWP,
CO2e
Carbon Dioxide
Carbon Monoxide
Methane
Nitrous Oxide

HTP,
1,4-DCBe*

ETP,
1,4-DCBe*

1
1.53
25
298

Carbon Tetrachloride

1,400

220

0.000143

Sulfur Hexafluoride

22,800

Benzene

1,900

6.35E5

Toluene

0.327

5.04E5

Xylene

0.43

0.382

S HTP

S ETP

Impact Indicator for


Impact Category

S GWP

* HTP and ETP are expressed in units of kg 1,4-dichlorobenzene equivalent (1,4-DCBe)

32

www.aiche.org/cep October 2015 CEP

LCA process nevertheless improves understanding of the


environmental and health impacts associated with each
alternative, where the impacts occur (locally, regionally,
or globally), and the relative magnitude of each type of
impact in comparison to each of the other proposed alternatives considered. This information more fully reveals the
pros and cons of each alternative.
Normalization and weighting. Because it is rare to find
an alternative that performs best in all impact categories,
comparative LCAs employ normalization and weighting.
Normalization converts the units of the impact categories
into a single dimensionless unit. Weighting (or valuation)
reflects the relative importance of the normalized environmental impacts according to the stakeholders and the
decision-makers preferences and values.
Normalization can be done internally or externally, or
both. Internal normalization utilizes values within the study
and shows the relative significance of impacts of competing alternatives (Figure 7).
External normalization uses an absolute scale based on
a reference system and requires an external database (13).
Normalized values are calculated by:
Ni = Si/Ai
where Ni is the normalized value for impact category i,
S is the characterized impact, and A is the normalization
reference value, such as a regions total emissions. Table 4
(13) presents the normalization factors for the U.S. that
the model TRACI 2.1 uses. For example, greenhouse
gas emissions from a product system can be normalized
relative to total global warming emissions for the U.S. of
7.40E+12 CO2e.
100
Percent of Maximum

Table 3 shows sample characterization factors for


emissions that contribute to the impact categories of GWP,
human health toxicity potential (HTP), and ecological
toxicity potential (ETP). Similar factors can be developed
for other impact categories.

80
60
40
20
0

GWP

EUP

POCP

ACP

HTP

FF

 Product A
 Product B
GWP = Global Warming Potential
EUP = Eutrophication Potential
POCP = Photochemical Oxidation Formation Potential (Smog)
ACP = Acidification Potential
HTP = Human Health Toxicity Potential
FF = Fossil Fuels

p Figure 7. Internal normalization allows the comparison of these two


hypothetical products.
Copyright 2015 American Institute of Chemical Engineers (AIChE)

External normalization references introduce additional uncertainty to the LCA study because of the lack
of consensus in data. Deciding on appropriate normalization methods is still an area of controversy and continued
research (1).
Weighting converts the impact scores (or normalized
impact scores) into a single number:
W = (WFc Ic)

ized databases will be developed, new impact assessment


methods will be designed, and methods for uncertainty
analysis will be improved.
Software and databases. Although LCA practice requires a high degree of expertise and knowledge, the availability of sophisticated LCA software
(such as SimaPro and GaBi) and the ecoinvent and
U.S. LCI Commons has put LCA within reach of a much
wider user base. A large leap forward in usability has
been in the creation of a fully transparent, open-source
software platform called openLCA. Available since 2006,
the software and its source code are freely available for
conducting professional-quality LCA. The availability of
these tools will help to make life cycle environmental performance as predominant as safety and quality are today
in the design and development of products, technologies,
and services.
Life cycle sustainability assessment (LCSA). With
LCAs moving us beyond examining a single impact or life
cycle stage as the sole criterion for environmental goodness
(or badness), we have gained much knowledge through the
thousands of LCA studies that have been completed over
the past 20 years. The movement now is toward integrating
LCA into the LCSA framework.
Klpffer (14) represents this integration with life cycle
costing (LCC) and social life cycle assessment (SLCA) as:

where W is the weighted result for the impact category c,


WFc is the weighting factor for this impact category, and Ic
is the impact score for this impact category. If all the environmental impacts are considered to be equally important,
they are given the same weight. At the extreme, viewing a
single issue, such as global warming, as the only impact of
interest (importance) leads to placing 100% of the weight
on GWP and, in effect, ignoring (assigning a value of zero
to) all other impacts.
Although weighting allows for impacts to be aggregated into a single score for easier evaluation according to
appropriate preferences, weights are inherently subjective
and can vary depending on culture, political views, gender,
demographics, and the professional opinions of stakeholders. Consequently, single-score results are criticized by
some practitioners.
An iterative process. Interpreting LCA findings
involves comparing the data and results with previous findLCSA = LCA + LCC + SLCA
ings, and putting them in the proper context of decisionmaking and limitations. The iterative nature of the ISO
Organizational LCA. LCA is also moving beyond prodframework reflects this.
ucts and is being promoted within a broader organizational
If the uncertainties are too high, additional data may be
scope, mainly in the assessment of individual environneeded. If the sensitivity analysis shows that some decimental impacts such as GHG emissions and water. Those
sions are crucial, additional, more-refined analysis may be
experiences laid the groundwork for LCA of organizations.
needed. It is especially important to determine whether the
The benefits and the potential lessons from the life cycle
results of the impact assessment or the underlying invenperspective are not limited to products.
tory data are incomplete or unacceptable for drawing con Nonetheless, the assessment for an organization is often
clusions and making recommendations. If so, the phases of
even more complex than that for products. Most organizathe LCA framework must be repeated until the results can
tions are engaged in many product life cycles, so there are
support the original goal of
the study.
Table 4. The TRACI 2.1 model uses these normalization factors for U.S. impacts (13).

Future directions
LCA has evolved from a
specialty field practiced by a
handful of practitioners with
closely guarded databases,
to a widely used tool with
emphasis on transparency and
data sharing. LCA practice
will continue to develop in
many directions. Regional-

Global Warming

Impact per
year

Population

Impact per
person-year

Units

7.40E+12

3.08E+08

2.40E+04

CO2 equivalents

Ozone Depletion

4.90E+07

3.06E+08

1.60E-01

CFC-11 equivalents

Acidification

2.80E+10

3.08E+08

9.10E+01

SO2 equivalents

Eutrophication

6.60E+09

3.00E+08

2.20E+01

PM-2.5 equivalents

Smog Formation

4.20E+11

3.00E+08

1.40E+03

O3 equivalents

Respiratory Effects

7.40E+09

3.08E+08

2.40E+01

PM-2.5 equivalents

Fossil Fuel Depletion

5.30E+12

3.12E+08

1.70E+04

MJ surplus

Copyright 2015 American Institute of Chemical Engineers (AIChE)

CEP October 2015 www.aiche.org/cep

33

Environmental Management

Organization
Raw
Materials

Facility X
Production

Use

Product A

End of Life

t Figure 8. An organizational LCA can include


the entire organization, one brand made in multiple
facilities, or a single facility. Source: Adapted
from (15).

tive (www.lifecycleinitiative.org). The


sidebar on the next page lists some of its
many publications.
Product C
Raw
The founders of the Forum for SusProduction
Production
Use
End
of
Life
Materials
tainability through Life Cycle Innovation
(FSLCI) aim to create the first organizaOther
Other
Activities*
Activities*
tion for the broad life cycle community
* Other activities include business
travel, employee commuting, etc.
people with common interest in creating
and using life cycle information. Bringing
multiple product life cycles to study (Figure 8) (15). Many
together stakeholders, initiatives, and activities around the
departments and business divisions may be involved, and a
world under one umbrella will allow linkage of activities
large part of the environmental impacts can reside outside
and allow members to be more effective by building on each
the organizations gate, across the value chain.
others strengths. FSLCI has an active board of directors and
The UNEP/SETAC Life Cycle Initiative developed
an executive committee. Individuals can apply for memberCEP
a guidance document for Organizational LCA (15). The
ship through its website, www.fslci.org.
ISOs Technical Specification (TS) 14072 provides guidance to enable organizations to more easily and more
effectively apply ISO 14040 and ISO 14044 at the organiAdditional Resources*
zational level (16).
European Platform on Life Cycle Assessment
http://eplca.jrc.ec.europa.eu
International efforts for LCA advancement
Raw
Materials

Facility Y

Production

Production

Use

Product B

The European Platform on Life Cycle Assessment


(http://eplca.jrc.ec.europa.eu) is a project of the European
Commission. It is carried out by the Commissions Joint
Research Centres (JRC) Institute for Environment and
Sustainability (IES), in collaboration with DG Environment,
Directorate Green Economy. Some of its publications are
listed in the sidebar at the right.
The United Nations Environment Programme (UNEP)
and the Society for Environmental Toxicology and Chemistry (SETAC) joined forces to advance life cycle thinking
worldwide through the UNEP/SETAC Life Cycle Initia-

MARY ANN CURRAN, PhD, is an internationally recognized expert in the field


of life cycle assessment (LCA) and sustainability, and is a co-owner
of BAMAC, Ltd. (Rock Hill, SC 29730; Phone: (803) 324-9551; Email:
macurran5137@gmail.com), which provides consulting services to clients
conducting LCAs. Until 2013, she worked in the U.S. Environmental
Protection Agencys Office of Research Development, where her activities
included developing LCA methodology, promoting LCA within the agency,
reviewing life cycle case studies, planning life cycle workshops and
conferences, and developing life cycle data and resources. She currently
serves as editor-in-chief of Springers International Journal of Life Cycle
Assessment. She has authored numerous papers and book chapters
that address the LCA concept and its applications, including Life Cycle
Assessment Handbook (2012), Life Cycle Assessment Student Handbook
(2015), and the LCA Compendium The Complete World of Life Cycle
Assessment book series (beginning in 2014). Curran holds a PhD in
cleaner production and sustainability from Erasmus Univ. in the Netherlands, an MSc in environmental management and policy from Lund Univ.
in Sweden, and a BS in chemical engineering from the Univ. of Cincinnati
in the U.S. She is a Fellow of AIChE.

34

www.aiche.org/cep October 2015 CEP

End of Life

International Reference Life Cycle Data System (ILCD)


Handbook, EUR 24982 EN; 72 pages (2012).
General Guide for Life Cycle Assessment Detailed
Guidance, EUR 24708 EN; 417 pages (2010).
General Guide for Life Cycle Assessment Provisions
and Action Steps, EUR 24378 EN, 163 pages (2010).
Specific Guide for Life Cycle Inventory Data Sets,
EUR 24709 EN, 142 pages (2010).
Analysis of Existing Environmental Impact Assessment
Methodologies for Use in Life Cycle Assessment, Background Document; 115 pages (2010).
Framework and Requirements for Life Cycle Impact
Assessment Models and Indicators, EUR 24586 EN,
116 pages (2010) .
Recommendations for Life Cycle Impact Assessment
in the European Context Based on Existing Environmental Impact Assessment Models and Factors,
EUR 24571 EN, 159 pages (2011).
Review Schemes for Life Cycle Assessment,
EUR 24710 EN, 34 pages (2010).
Reviewer Qualification for Life Cycle Inventory Data Sets,
EUR 24379 EN, 34 pages (2010).
* All publications are in English unless otherwise noted.

Copyright 2015 American Institute of Chemical Engineers (AIChE)

Additional Resources*
UNEP/SETAC Life Cycle Initiative
www.lifecycleinitiative.org
Guidance on Organizational Life Cycle Assessment,
148 pages (2015).
Life Cycle Thinking in Latin America, 16 pages (2015).
Hotspots Analysis: Mapping of Existing Methodologies,
Tools, and Guidance and Initial Recommendations for the
Development of Global Guidance, 179 pages (2014).

Literature Cited
1. Curran, M. A., Student Life Cycle Assessment Handbook,
Scrivener-Wiley Publishing, Salem, MA (2015).
2. International Organization for Standardization, Environmental Management Life Cycle Assessment Principles and
Framework, ISO 14040, ISO, Geneva, Switzerland (2006).
3. International Organization for Standardization, Environmental Management Life Cycle Assessment Requirements
and Guidelines, ISO 14044, ISO, Geneva, Switzerland (2006).
4. Finnveden, G., et al., Recent Developments in Life Cycle
Assessment, Journal of Environmental Management, 91 (1),
pp. 121 (2009).

An Analysis of Life Cycle Assessment in Packaging for


Food and Beverage Applications, 81 pages (2013).

5. Sjunnesson, J., Life Cycle Assessment of Concrete, Masters


Thesis, Lund Univ., Lund, Sweden (2005).

The Methodological Sheets for Sub-Categories in Social


Life Cycle Assessment (S-LCA), 144 pages (2013).

6. Shapouri, H., et al., The Energy Balance of Corn Ethanol: An


Update, Agricultural Economic Report No. 813, U.S. Dept. of
Agriculture, Office of the Chief Economist, Office of Energy
Policy and New Uses, Washington, DC (2002).

Greening the Economy through Life Cycle Thinking,


60 pages (2012).
Towards a Life Cycle Sustainability Assessment,
86 pages (2011).

7. Curran, M. A., Sourcing Life Cycle Inventory Data, Chapter 5


in Life Cycle Assessment Handbook: A Guide for Environmentally Sustainable Products, Curran, M.A., ed., Scrivener-Wiley
Publishing, Salem, MA (2012).

Global Guidance Principles for LCA Databases: A Basis


for Greener Processes and Products, 160 pages (2011).

8. Searchinger, T., et al., Use of U.S. Cropland for Biofuels


Increases Greenhouse Gases through Emissions from Land-Use
Change, Science, 319 (5867), pp. 12381240 (2008).

Guidelines for Social LCA of Products, 104 pages, available in English, French, and Dutch (2009).

9. Hertwich, E. G., et al., Introduction, Chapter 1 in Life Cycle


Impact Assessment Striving Towards Best Practice, Udo de
Haes, H. A., et al., eds., SETAC Press, Brussels, Belgium (2002).

LCM: How Business Uses it to Decrease Footprint,


Create Opportunities, and Make Value Chains More
Sustainable, 48 pages (2009).
Guidance on How to Move from Current Practice to
Recommended Practice in Life Cycle Impact Assessment,
33 pages (2008).
Communication of Life Cycle Information in the Building
and Energy Sectors, 93 pages (2008).
Life Cycle Management: A Business Guide to Sustain
ability, 52 pages (2007).
Background Report for a UNEP Guide to Life Cycle
Management, 108 pages (2006).
Life Cycle Approaches: The Road from Analysis to
Practice, 89 pages (2005).
Why Take a Life Cycle Approach, 28 pages, available in
English, Chinese, Spanish, French, and Japanese (2004).
Report of the LCM Definition Study Peer Review,
79 pages (2003).
Life Cycle Impact Assessment Methods, 25 pages (2003).
Evaluation of Environmental Impacts in Life Cycle
Assessment, 108 pages (2003).
* All publications are in English unless otherwise noted.

Copyright 2015 American Institute of Chemical Engineers (AIChE)

10. European Commission Joint Research Centre, Framework


and Requirements for Life Cycle Impact Assessment Models
and Indicators, in ILCD Handbook International Reference
Life Cycle Data System, EC JRC, Institute for Environment and
Sustainability, Ispra, Italy (2010).
11. Bare, J. C., and T. P. Gloria, Critical Analysis of the Mathematical Relationships and Comprehensiveness of Life Cycle
Impact Assessment Approaches, Environmental Science and
Technology, 40 (4), pp. 11041113 (2006).
12. Jolliet, O., et al., Global Guidance on Environmental Life
Cycle Impact Assessment Indicators: Findings of the Scoping
Phase, International Journal of Life Cycle Assessment, 19 (4),
pp. 962967 (2014).
13. Ryberg, M., et al., Updated U.S. and Canadian Normalization
Factors for TRACI 2.1, Clean Technology and Environmental
Policy, 16 (2), pp. 329339 (2014).
14. Klpffer, W., Introducing Life Cycle Assessment and its Presentation in LCA Compendium, Chapter 1 in LCA Compendium:
The Complete World of Life Cycle Assessment, Klpffer, W.,
and M. A. Curran, eds., Springer Dordrecht, Germany (2014).
15. United Nations Environment Programme and Society for
Environmental Toxicology and Chemistry, Guidance on
Organizational Life Cycle Assessment, UNEP/SETAC, Paris,
France (2015).
16. International Organization for Standardization, Environmental Management Life Cycle Assessment Requirements and Guidelines for Organizational Life Cycle Assessment, ISO Technical Specification (TS) 14072, ISO, Geneva,
Switzerland (2014).

CEP October 2015 www.aiche.org/cep

35

Potrebbero piacerti anche