Sensory Shelf Life Estimation A Review o

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Food Research International 49 (2012) 311–325

Contents lists available at SciVerse ScienceDirect

Food Research International


journal homepage: www.elsevier.com/locate/foodres

Review

Sensory shelf-life estimation: A review of current methodological approaches


Ana Giménez ⁎, Florencia Ares, Gastón Ares
Departamento de Ciencia y Tecnología de Alimentos, Facultad de Química, Universidad de la República, General Flores 2124, C.P. 11800, Montevideo, Uruguay

a r t i c l e i n f o a b s t r a c t

Article history: Shelf-life of food products can be regarded as the period of time during which a product could be stored until
Received 30 May 2012 it becomes unacceptable from safety, nutritional, or sensory perspectives. Shelf-life estimation of food prod-
Accepted 6 July 2012 ucts and beverages has become increasingly important in recent years due to technological developments
and the increase in consumer interest in eating fresh, safe and high quality products. The shelf-life of the ma-
Keywords:
jority of food products is determined by changes in their sensory characteristics. Therefore, in order to extend
Shelf-life
Consumer studies
commercialization times to its maximum while assuring products' quality, food companies should rely on ac-
Sensory evaluation curate methodologies for sensory shelf-life estimation. Despite several methodologies have been developed
Consumer research in the last decade, their application in the mainstream food science and technology literature is still limited
Survival analysis and most studies dealing with sensory shelf-life rely on basic and inaccurate approaches. In this context,
Liking the aim of this work is to review current methodological approaches for sensory shelf-life estimation. Imple-
mentation, applications, advantages and disadvantages of quality-based methods, acceptability limit, cut-off
point methodology and survival analysis are discussed. The superiority of consumer-based methodologies is
highlighted, with the aim of encouraging researchers to base their sensory shelf-life estimations on consumer
perception.
© 2012 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
312
1.1. Sensory shelf-life: concept and relevance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
312
2. Design of sensory shelf-life experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312
2.1. Basic design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
2.2. Reversed design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
3. Methodologies for sensory shelf-life estimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
3.1. Quality-based methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
3.1.1. Difference from control test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
3.1.2. Intensity of sensory attributes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
3.1.3. Quality rating methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316
3.2. Acceptability limit methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
3.3. Cut-off point methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
3.4. Survival analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
3.4.1. Current status survival analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
4. Methodological recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
5. Challenges and suggestions for further research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323

⁎ Corresponding author. Tel.: +598 2924 80 03; fax: +598 2924 19 06.
E-mail address: agimenez@fq.edu.uy (A. Giménez).

0963-9969/$ – see front matter © 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.foodres.2012.07.008
312 A. Giménez et al. / Food Research International 49 (2012) 311–325

1. Introduction technologies request stability testing to assure foods are safe and
have an acceptable quality when consumed.
1.1. Sensory shelf-life: concept and relevance In this context, estimating shelf-life of food products and bever-
ages has become increasingly important in recent years (Stone &
Shelf-life is usually defined as the time during which a food product Sidel, 2004). A quick search in Scopus database reveals a clear in-
remain safe, comply with label declaration of nutritional data and retain crease in the number of articles published in peer reviewed interna-
desired sensory, chemical, physical and microbiological characteristics tional journals which included the words shelf-life and food in their
when stored under the recommended conditions (IFST, 1993). There- title, abstract or keywords. As shown in Fig. 1, the number of articles
fore, to assess shelf-life objective indexes related to nutrition, microbio- has increased from 532 in 2002 to 1579 in 2011.
logical or physicochemical characteristics of food have been typically According to Dethmers (1979), both analytical as well as affective sen-
measured (Cardello, 1995; Wansink & Wright, 2006). sory methods may be used to determine shelf-life of food and actually
Shelf-life is a function of time, environmental factors, and suscepti- complement each other. Regardless of the method selected and the ratio-
bility of product to quality change (Labuza & Szybist, 2001). Physical, nale behind, sensory evaluation is a key factor for determining the
chemical and biological changes that occur throughout the food chain shelf-life in several food categories. Being sensory shelf-life dependant
generally lead to product deterioration and these changes might in on consumers' judgment of whether a food product is acceptable or not,
time compromise nutritional, microbiological or sensory quality. In it is essential that results from any instrumental or chemical analysis cor-
many products changes in sensory characteristics occur largely before relate closely with results from sensory evaluation (Robertson, 2006).
any risk to consumers' health is reached (Lawless & Heymann, 2010). Griffiths (1985) found that significant changes in descriptive ratings do
According to Hough (2010) the shelf-life of most food products is limit- not always translate to significant differences in consumer acceptability,
ed by changes in their sensory characteristics. In this context, sensory highlighting the importance of consumer perspective. Even though con-
shelf-life estimation of foods has become an issue of continuous and ex- sumer acceptability is of key importance, knowing the sensory changes
tensive research on both the deteriorative mechanisms occurring in that occurred and how these changes affected acceptability would pro-
food systems and the development and application of methodologies vide valuable information for manufacturers. Identifying the sensory
for shelf-life estimation (Manzocco & Lagazio, 2009). factor that limits the sensory shelf-life of a food product could help man-
Accurate prediction of shelf-life is essential to consumers as well ufacturers to select formulation or processing conditions that improve
as manufacturers. Consumer increasing concerns on healthy eating product quality throughout storage. Martínez, Ares, and Lema (2008)
and making healthy food choices demand among others, fresh, conve- reported that the sensory shelf-life of fresh-cut butterhead lettuce was
nient, safe and superior quality foods. A growing concern over wheth- limited by browning, suggesting that the application of antioxidants
er the food they purchase is fresh or not, or how long it will keep its that minimize this sensory defect could positively contribute to improv-
quality is one of the reasons driving consumers to read labels. ing the product. Conte, Brescia, and Del Nobile (2011) reported that the
Shelf-life dating is considered by most consumers to be a measure sensory shelf-life of Burrata cheese was determined by changes in consis-
of food freshness, relying on the information provided by the manu- tency, which led them to select lysozyme/Na2-EDTA with modified atmo-
facturer when making their purchase decisions (OTA, 1979). The sphere packaging for shelf-life extension. Similarly, Jacobo-Velázquez and
magnitude of perceived risk (Murray & Schlacter, 1990; Severson, Hernández-Brenes (2011) concluded that sour flavor development dur-
Slovic, & Hampson, 1993) and previous experience with a product ing storage due to rupture of cell membranes and diffusion of intracellular
(Weber & Milliman, 1997) are likely conditions that influence con- organic acid should be improved in order to extend the sensory shelf-life
sumers to check product expiration dates. Several authors have of high hydrostatic pressure processed avocado paste.
reported that information presented on labels could have a major in- The aim of this work is to review the implementation of current
fluence on food acceptance (Jaeger, 2006; Rozin, 1990; Rozin & methodological approaches for sensory shelf-life estimation and to
Tuorila, 1993), which suggests that shelf-life dates could significantly discuss applications, advantages and disadvantages.
influence consumer expectations and perception of food products.
The economic impact of a business decision based in part on inappro- 2. Design of sensory shelf-life experiments
priate shelf-life dating can be significant (Stone & Sidel, 2004). Finding
unacceptable products within their shelf-life could diminish consumer Sensory shelf-life estimation of a food product basically consists
confidence in the brand and in the store that sells it, leading them to on the evaluation of the sensory characteristics of a set of samples
not purchasing that particular brand again (Harcar & Karakaya, 2005). with different storage times (Bishop & White, 1986).
On the other hand, the financial impact of retrieving an acceptable prod- The following steps could be identified when designing a sensory
uct from the marketplace should also be considered (Mena, Adenso-Diaz, shelf-life experiment (Dethmers, 1979; Peryam, 1964): (i) determining
& Yurt, 2011). It has been estimated that between 25% and 50% of food
production is wasted along the supply chain (Nellman et al., 2009).
Throwing food away has economic and environmental implications
(Stuart, 2009; Ventour, 2008), and has also raised moral questions consid-
ering the number of people who suffer from hunger worldwide (Stuart,
2009). Besides, accurate shelf-life labeling could contribute to an effective
waste management which could increase profitability levels all along the
supply chain, which is particularly relevant considering the low profitabil-
ity margins traditionally associated to food industry (Hyde, Smith, Smith
& Henningson, 2001). Therefore, in order to extend commercialization
times to its maximum while assuring the product's freshness, food com-
panies should rely on accurate methodologies for shelf-life estimation
(Giménez, Ares, & Gámbaro, 2008a).
During the last decade, technological developments as well as new
packaging materials have been developed as strategies for modern
food preservation in order to meet growing consumer demands for
safe and durable food products offering high nutritional and sensory Fig. 1. Number of articles included in Scopus database including the words shelf-life and
value (Walkling-Ribeiro, Noci, Cronin, Lyng, & Morgan, 2009). These food in their title, abstract or title from 2002 to 2011.
A. Giménez et al. / Food Research International 49 (2012) 311–325 313

the objectives of the study, (ii) getting representative samples from the fresh samples to the assessors in each evaluation could be an easy way
test product, (iii) determining the physical and chemical composition of of minimizing this type of bias.
the products, (iv) selecting the storage conditions, (v) setting up a test
design or defining how the samples are going to be stored and evaluat- 2.2. Reversed design
ed, (vi) selecting an appropriate methodology, (vii) establishing the
criteria that will be considered for defining the sensory shelf-life of Another option for designing a sensory shelf-life experiment is to
the product, (viii) conducting the experiment, and (ix) analyzing the re- evaluate a set of samples with different storage times, all together, at
sults and estimating the sensory shelf-life of the product. a single evaluation instance. This type of design is called reversed
The setting up of the sensory shelf-life experiment defines the storage design and has the advantages of overcoming the major
time and resources needed and therefore is one of the most important drawbacks of basic storage design (Hough, 2010).
points that should be taken into account. According to Robertson Reversed storage design can be performed by staggering product
(2006), an issue of shelf-life testing is developing an experimental de- times, so that all products with different storage times are evaluated on
sign that minimizes the cost and time of the testing while providing the same day (Lawless & Heymann, 2010). Gámbaro, Ares, and Giménez
reliable and statistically valid data. Two main strategies for storing (2006) followed this approach for estimating the sensory shelf-life of
and evaluating products during a sensory shelf-life experiment have apple-baby food by working with different industrial batches stored at
been used: basic and reversed storage design (Hough, 2010). 25.0±0.5 °C for 0, 7, 12, 24, 30, 33, 35, 40 and 46 months. As shown in
Fig. 3, fresh apple-baby food samples from different batches were placed
2.1. Basic design in the temperature-controlled room at several different times, so that at
the evaluation day samples with 9 different storage times were evaluated.
Basic storage design is the simplest and most common approach for In order to use this approach it is crucial to have homogeneous batches of
performing a sensory shelf-life experiment (Hough, 2010). It consists of products throughout a long period of time. Gámbaro, Ares, et al. (2006)
storing a single large batch of product under normal conditions and to reported that previous studies have proved that the industrial process
test it at various storage times (Lawless & Heymann, 2010). variation for apple-baby food was minimal and therefore differences in
Fernández-López et al. (2008) used this type of design for estimat- the sensory characteristics of the evaluated samples could be attributed
ing the shelf-life of ostrich. These authors stored all the ostrich steaks only to differences in storage time.
at 2 °C and removed samples after 0, 4, 8, 12 and 18 days of storage When it is not feasible to get homogeneous batches, reversed stor-
for sensory and physicochemical analysis (Fig. 2). age design could be implemented by storing the product under condi-
This type of design has been used for sensory shelf-life estimation tions that stop all deterioration processes, for example by freezing or
of a wide range of food products, including commercial mayonnaise storing at very low refrigeration temperatures (Lawless & Heymann,
(Martínez, Mucci, Santa Cruz, Hough, & Sánchez, 1998), European 2010). Giménez et al. (2007) followed this approach for estimating
hake (Rodríguez, Losada, Aubourg, & Barros-Velázquez, 2004) dark the shelf-life of brown pan bread. As shown in Fig. 4, these authors
chocolate (Nattress, Ziegler, Hollender, & Peterson, 2004), Fuji ap- stored breads in a temperature-controlled storage room at 20 °C for
ples (Varela, Salvador, & Fiszman, 2005), “flor de invierno” pears 1, 4, 7, 10, 13, 15 and 17 days. After reaching the desired storage
(Salvador, Varela, & Fiszman, 2007), butterhead lettuce (Lareo et times, breads were frozen at − 20 °C and stored at − 18 °C, providing
al., 2009), probiotic yogurt (Cruz et al., 2010), chocolate and carrot samples with different storage times from one batch. On the evalua-
cupcakes (Montes Villanueva & Trindade, 2010), and minimally tion day, samples were defrosted at 20 °C for 6 h before their sensory
processed kiwifruit (Mastromatteo, Conte, & Del Nobile, 2011). evaluation. This type of reversed storage design enabled the evalua-
Although basic storage design is the most common approach for tion of pan bread samples with different storage times on a single
sensory shelf-life experiments, it is not efficient in terms of use of evaluation day, which minimizes the time and resources needed for
time of resources (Lawless & Heymann, 2010). When a shelf-life exper- the experiment. Before using this type of design it was necessary to
iment is performed following a basic design, sensory and physicochem- verify that the frosting–defrosting cycle did not significantly modify
ical analysis should be performed at each storage time. In the example the sensory characteristics of pan bread.
depicted in Fig. 2, the sensory panel should evaluate ostrich samples A similar approach has been also used by Jacobo-Velázquez, Ramos-
in 5 different occasions. Additionally, if sensory shelf-life is estimated Parra, and Hernández-Brenes (2010) for estimating the sensory shelf-
using consumer data, performing 5 studies with 50–100 consumers life of high hydrostatic pressure processed avocado pulp at 4 °C.
would considerably increase the total cost of the experiment. A variation of the abovementioned procedure would be to first store
Another disadvantage of basic storage design is the risk that the the product under conditions that minimize deterioration, then to pull
trained assessor and/or consumer panel changes its criteria (Lawless products from those optimal conditions at different times and to store
& Heymann, 2010). Assessors can become aware of the aim of the ex- them under normal storage conditions to allow them to deteriorate
periment and expect that samples become more deteriorated as time prior to their evaluation (Lawless & Heymann, 2010). This approach
passes by, which could lead to biased results (Hough, 2010). Presenting was followed by Hough, Langohr, Gómez, and Curia (2003), who kept

Storage of all the


ostrich steaks at 2±1οC

Time (days)

0 4 8 12 16

Sensory and Sensory and Sensory and Sensory and Sensory and
physicochemical physicochemical physicochemical physicochemical physicochemical
analysis analysis analysis analysis analysis

Fig. 2. Example of a basic storage design for estimating the sensory shelf-life of ostrich steaks at 2 °C.
314 A. Giménez et al. / Food Research International 49 (2012) 311–325

Storage of fresh apple-baby food samples from different industrial batches at 25.0 ± 0.5οC

Time (months)

-46 -40 -35 -33 -30 -24 -12 -7 0

Sensory and
physicochemical
analysis

Fig. 3. Example of a reversed storage design for estimating the sensory shelf-life of apple-baby food at 20 °C using different industrial batches.

yogurt samples at 4 °C and periodically place them at 42 °C to get sam- noticeable difference, defined as the length of time necessary to distin-
ples with 0, 4, 8, 12, 24, 36 and 48 h at that temperature. guish the product being evaluated from the fresh product (Heldman &
The major advantage of reversed storage design is that all samples Hartel, 1997); (iv) significant differences in descriptive analysis profile
are evaluated at the same time, which minimizes the time, effort and from the fresh product; (v) correlating analytical and affective data
resources necessary for carrying out the experiment. This is particu- and establishing a time to failure (Gacula, 1975; Gacula & Kubala,
larly useful when using consumer studies for sensory shelf-life esti- 1975); (vi) predetermined consumer overall liking scores or rejection
mation. However, it is not always possible to have homogeneous to consume or purchase the product.
batches or to find storage conditions that minimize sensory changes. In this review, the most popular methodologies for sensory shelf-life
For example, it is not easy to find storage conditions that minimize estimation are presented: quality-based methods, acceptability limit,
the deterioration process of fresh and perishable products such as cut-off point methodology and survival analysis.
fruit or vegetables.
3.1. Quality-based methods
3. Methodologies for sensory shelf-life estimation
One of the most common approaches for estimating sensory shelf-
Once the strategy for storing and evaluating products has been se- life has been measuring quality throughout storage using a trained as-
lected, the methodology for monitoring the sensory quality of the sessors' panel or a group of experts. The sensory shelf-life of the product
products and for estimating the storage period that corresponds to is defined as the storage time at which overall quality or the intensity of
its sensory shelf-life should be defined. As highlighted by Lawless a specific sensory attribute reaches a predetermined value or “failure
and Heymann (2010), sensory shelf-life testing is the repeated appli- criterion”, assuming that once the product has reached this point it is
cation of common sensory methodologies rather than a special senso- not longer salable (Lawless & Heymann, 2010). Within this main ap-
ry methodology. Depending on the specific aim of the study, sensory proach, three different measurements have been considered: overall
shelf-life experiments could be performed by applying discrimina- difference with respect to the fresh sample, overall quality, and attri-
tion, descriptive or affective methodologies (Kilcast, 2000). bute intensity.
No matter what methodology is selected, sensory shelf-life esti-
mation requires the selection of failure criteria or a cut-off point, 3.1.1. Difference from control test
which corresponds to the maximum deterioration that is considered Sensory shelf-life can be estimated by measuring the degree of differ-
acceptable. In other words, sensory shelf-life is estimated as the stor- ence between stored samples and a fresh product, considered as control.
age time when the product reaches a certain predetermined deterio- Specifically, a sensory panel is trained in measuring the degree of differ-
ration level, above which it is not salable. Several failure criteria have ence between products using discriminative tests or intensity scales
been used when considering sensory data (Dethmers, 1979): (i) fixed (Lawless & Heymann, 2010). Once the panel is trained, sensory tests
increase or decrease in the average intensity of a sensory attribute are conducted to determine the degree of difference between stored
(Gacula, 1975; Gacula & Kubala, 1975); (ii) storage life, defined as samples and a fresh control sample. The magnitude of the difference be-
the time needed to reach an undesired overall quality; (iii) just tween stored samples and the control is regressed as a function of

Storage of all the brown


pan breads at 20οC

Time (days) at 20οC

0 1 4 7 10 13 15 17

Storage at -18οC

Time (days) at -18οC

0 1 4 7 10 13 15 17
Defrosting

Sensory analysis

Fig. 4. Example of a reversed storage design for estimating the sensory shelf-life of brown pan bread at 20 °C by storing the samples at −18 °C.
A. Giménez et al. / Food Research International 49 (2012) 311–325 315

storage time and sensory shelf-life is estimated as the time period at A second alternative is to measure the degree of difference be-
which the product reaches a predetermined difference from the fresh tween stored samples and a fresh control using scales. Hough et al.
control product. (1999) used a difference-from-control test to determine the sensory
One of the key points of the methodology is finding the storage con- shelf-life of ricotta cheese at 6 °C. These authors asked a trained as-
ditions for the fresh control sample, so that it is available and unaltered sessor panel to measure the degree of the difference between the
throughout the whole storage period considered in the experiment. For stored samples and a fresh control using a 7-point structured scale
some types of products it is relatively simple to find storage conditions (0 = no difference, 6 = very big). In this methodology the assessors
that minimize their deterioration process. In the case of frozen food are usually trained by evaluating a control and samples with different
products stored at −18 °C, the control is usually a product stored at − storage times. By open discussion with the panel leader the assessors
30 °C (Symons, 2000). Similarly, Patsias, Chouliara, Badeka, Savvaidis, agree on the score which corresponds to the difference between each
and Kontominas (2006) considered a freshly thawed chicken sample sample and the control. During the shelf-life study, in each evaluation
stored at −30 °C throughout the experiment as control when studying the trained assessors received the control sample labeled as K, two
the influence of modified atmosphere packaging on the shelf-life of pre- stored samples and a blind control coded with 3-digit numbers. If
cooked chicken product stored under refrigeration. Besides, when esti- trained assessors are asked to evaluate a blind control sample using
mating the sensory shelf-life of mayonnaise at 20–45 °C, Martínez et al. difference-from-control scale they usually provide a value different
(1998) stored the control at 5 °C since sensory changes at this tempera- from zero, which makes it necessary to correct the scores. Therefore,
ture were regarded as insignificant compared to those occurring at the in this procedure assessors usually evaluate a blind control sample
temperatures considered in the shelf-life study. Before selecting the stor- and the sample score is obtained by subtracting the blind control's av-
age conditions for the control, it is necessary to assure that they do not erage score from the sample's average (Meilgaard, Civille, & Carr,
cause any significant changes in their sensory characteristics. A triangle 1991). In order to estimate sensory shelf-life the authors regressed
test with a trained panel should be performed to compare fresh and the average difference between the stored ricotta cheese and the
stored samples. In this test three samples are presented simultaneously fresh control and defined sensory shelf-life as the time at which the
to the assessors, two of which are identical and one is different. Each as- degree of the sensory difference between the stored samples and
sessor has to indicate which is the odd sample, and binomial tests are the fresh control reached an average score of 1.5 (corresponding to
used to determine if significant differences exist between a fresh sample a description between very slight and slightly different). This criteri-
and one stored at the selected conditions (Lawless & Heymann, 2010). on was selected considering that consumers are not likely to tolerate
When it is not possible to keep the same control during the whole changes in the sensory characteristics of ricotta cheese.
experiment some authors have periodically replaced it for a new fresh Freitas and Costa (2006) followed a similar approach with a
sample. For example, when working with ricotta cheese stored at manufactured dehydrated product and asked a trained assessor
6 °C, Hough, Puglieso, Sánchez, and Mendes Da Silva (1999) stored panel to rate the degree of difference between stored products and
the control at 2 to 3 °C. This sample maintained its sensory characteris- a reference using a 7-point scale (0–6), considering that products
tics unchanged for a period of only 7 days. Therefore, after 7 days the with a score of 3 were unfit for human consumption.
control was replaced by a new fresh sample and triangle tests with a Martínez et al. (1998) proposed a variation of this method for es-
trained panel were carried out to ensure that they were not significantly timating the sensory shelf-life of mayonnaise. At each storage time,
different. Following a similar approach, Alkadamany et al. (2002) re- trained assessors evaluated the degree of difference between the
placed control samples of concentrated yogurt (labneh) stored at 5 °C stored samples and a fresh control in 9 specific sensory attributes
every 3 days for estimating sensory shelf-life at 5, 15 and 25 °C. (total aroma, acid, salty, lemon, egg, oily, oxidized, heat and isothio-
The degree of difference between the stored samples and the fresh cyanate), instead of considering global differences. For each descrip-
control could be estimated using two main methodological approaches: tor a 12 cm unstructured scale was used, anchored at the center
discriminative tests or intensity scales. with “equal to control”, at the extreme left with “a lot less than con-
In the first approach the trained assessor panel (composed of at least trol”, and at the extreme right with “a lot more than control”. A linear
10 assessors) carry out paired comparisons, triangle or a duo–trio test in regression between difference with control and storage time was
order to determine whether the stored sample and a control fresh sam- used to estimate sensory shelf-life. The failure criterion was a differ-
ple are perceptibly different (Lawless & Heymann, 2010). Sensory ence of ±1.5 cm in the 12 cm scale.
shelf-life is defined as the length of time during which the product The advantage of this method is that it provides an accurate estima-
does not significantly change its sensory characteristics, which corre- tion of the time at which changes in the sensory characteristics of the
sponds to the high quality life (HQL) (Heldman & Hartel, 1997). This ap- samples occur. However, it does not necessarily mean that detecting
proach has been used for estimating the sensory shelf-life of frozen differences between samples with different storage times will lead to
foods, by performing triangle tests between the samples stored at − rejection of the product by consumers. Therefore, shelf-life estimations
18 °C and a control sample stored at −30 °C (Symons, 2000). Using bi- obtained from this methodology are usually too conservative and could
nomial tests, the sensory shelf-life is estimated as the time at which the lead to a decrease in commercialization times and profitability levels for
first significant difference between the stored samples and the control is the manufacturer.
detected, which corresponds to the first just noticeable or perceivable
change in the sensory characteristics of the product.
Paired comparisons could also be used instead of triangle tests for sen- 3.1.2. Intensity of sensory attributes
sory shelf-life estimations. Paired comparisons are discriminative tests Another popular approach for sensory shelf-life estimation has been
used when the experimenter wants to determine if two samples differ measuring the intensity of sensory attributes throughout storage and
in a specified sensory characteristic. Two samples are simultaneously estimating shelf-life as the time at which the intensity of a critical attri-
presented to the assessor, who has to identify the sample that is higher bute reaches a certain predetermined value. In this approach trained as-
in the specified sensory attribute (Lawless & Heymann, 2010). This ap- sessors are asked to try the product and then generate a numerical
proach has been followed by Schmidt and Bouma (1992) to estimate response using a scale that reflects how they perceived the intensity
the sensory shelf-life of cottage cheese at 4 and 7 °C, considering samples of one or more of the sensations generated by that product (Lawless &
stored at 0 °C as control. At each storage time, the assessors were Heymann, 2010). Different types of scales could be used for this pur-
presented with paired comparisons between the stored samples and pose, being unstructured scales widely recommended (Stone & Sidel,
the control, and were asked to identify which sample had more “fruity 2004). When using unstructured scales assessors are asked to indicate
fermented off-flavor”. the intensity of the sensory attribute by making on a 10 or 15 cm line.
316 A. Giménez et al. / Food Research International 49 (2012) 311–325

The key step of this methodology is how to select the attributes that Although a wide range of methods for measuring sensory quality
are responsible for the sensory changes throughout storage. A common have been developed (Muñoz, Civille, & Carr, 1992), the quality rating
approach in order to identify the defects most likely to appear due to method is by far the most popular for sensory shelf-life estimation.
prolonged storage, is to perform a descriptor generation step by consen- This method consists in asking a trained assessor panel to rate the
sus in an open session. Assessors are provided with a set of samples with quality of products using a scale in which points are defined in
different storage times and are asked to generate the descriptors needed terms of the sensory characteristics that characterize the quality of
to describe differences among samples, by coming to some consensus each grade (Costell, 2002). It is important to take into account that
through open discussion with the panel leader (Lawless & Heymann, this method requires a highly trained sensory panel since assessors
2010). For example, in order to identify the sensory attributes for esti- should be able to use three main abilities: remembering the sensory
mating the sensory shelf-life of lettuce based on visual appearance, characteristics of the ideal product, interpreting the descriptions
Lareo et al. (2009) presented four samples of lettuce with different stor- corresponding to each point of the scale, identifying the common sen-
age times at 15 °C (0, 7, 10, and 17 days) and asked assessors to write sory defects that appear as a result of prolonged storage, and using
down the descriptors that made the appearance of those samples differ- the quality scale to quantify the level of severity of each defect
ent. Through open discussion with the panel leader, the assessors agreed (Lawless & Heymann, 2010).
on the appearance descriptors that best differentiated the stored sam- The Quality Index Method (QIM) is a good example of the applica-
ples and how to evaluate them. After the descriptors are selected, a tion of quality rating methods for sensory shelf-life estimation. This
training step on measuring attribute intensity should be performed by method is based on the objective evaluation of the key sensory attri-
subsequently exposing assessors to samples with different intensities of butes of each fish species using a scoring system that ranges from 0 to
each attribute and potential references standards (Lawless & Heymann, 3 (the lower the score the fresher the fish) (Costell, 2002). Cardenas
2010). Usually 10 or 15-cm unstructured line scales are used for rating at- Bonilla, Sveinsdottir, and Martinsdottir (2007) reported the develop-
tribute intensity and the assessors are first asked to select the words ment of a QIM scheme for fresh cod fillets and its application in a
needed to anchor the scales (such as none to very strong) and the shelf-life study. In order to develop a preliminary QIM scheme, two
reference standards which correspond to the lowest and highest in- researchers observed and registered the sensory changes of the cod
tensity of each attribute. A first session in which assessors discuss fillets from the day of filleting until spoiled. Each parameter was
and agree on the consensus intensity of the attributes could be car- rated using a 3-point scale that ranged from 0 (fresh) to 3 (spoiled).
ried out (Jacobo-Velázquez & Hernández-Brenes, 2011). In subse- Then, 11–12 assessors with previous experience in sensory evalua-
quent sessions assessors are asked to evaluate a set of samples and to tion were trained in the evaluation of each sensory parameter by
rate the intensity of the sensory attributes in several sessions, according evaluating cod fillets that differed in storage time. Through discussion
to the consensus or expected intensity score. Once the assessors provide with the panel leader, the assessors agreed on the final QIM scheme,
reproducible and consistent results, real samples could be evaluated dur- which included the evaluation of 8 sensory attributes (Table 1). The
ing the shelf-life study. sum of the scores of the 8 attributes corresponded to the Quality
The sensory shelf-life of different products has been estimated by Index (QI), which ranged from 0 for a fresh fillet to 18 to a spoiled fil-
measuring the intensity of specific sensory attributes. Nattress et al. let. This QIM scheme was used to evaluate the sensory shelf-life of
(2004) used this methodology for estimating the influence of hazelnut cod fillets stored at 1 °C on ice on plastic boxes after 0, 3, 7, 10 and
paste on the sensory shelf-life of dark chocolate. Throughout storage, a 14 days. For each evaluation instance, the average QI of the trained
panel of 10 trained assessors was asked to measure the intensity of ran- panel was calculated and linearly regressed as a function of storage
cid flavor using 15-cm unstructured scales. The sensory shelf-life of dark time. The sensory shelf-life of the cod fillets could be estimated by in-
chocolate containing hazelnut paste was estimated as the onset of rancid terpolating in the graph the time corresponding to a predetermined
flavor. Meanwhile, for evaluating the sensory shelf-life of whole milk QI value. However, the authors did not select a maximum QI.
powder, Nielsen, Stapelfeldt, and Skibsted (1997) asked between 9 and The application of QIM is the most common approach for estimat-
12 trained assessors to evaluate the intensity of oxidized flavor using a ing the sensory-shelf-life of fresh fish and seafood and has been ex-
16-point structured scale (0 = extremely oxidized flavor, 15 = no oxi- tended to 12 species (Costell, 2002), including striped red mullets
dized flavor). Scores equal to or higher than 10 were considered as ac- (Bono & Badalucco, 2012), blackspot seabream (Silva Sant'Ana,
ceptable, whereas lower scores indicated defective samples. Another Soares, & Vaz-Pires, 2011), cuttlefish and squid (Vaz-Pires et al.,
example of sensory shelf-life estimation by measuring attribute intensi- 2008) and European hake (Rodríguez et al., 2004).
ty was reported by Piagentini, Mendez, Guemes, and Pirovani (2005), Despite the popularity of the QIM method, other quality-based ap-
who asked a panel of trained assessors to evaluate off-odor, general proaches have also been applied for estimating the sensory shelf-life of
appearance, wilting, and browning of fresh-cut lettuce using a 15 cm sea products. Jeevanandam, Kakatkar, Doke, Bongirwar, and Venugopal
unstructured scale. These authors considered the fresh-cut lettuce as (2001) estimated the sensory shelf-life of threadfin bream (Nemipterus
unacceptable when a mean score below 7.5 was reached for general ap- japonicus) in ice using a 10-point quality scale for odor. Fish that scored
pearance or above 7.5 for the other sensory attributes. between 10 and 5 were considered acceptable and those scoring less
Other authors have used structured scales with a reduced number of than 5 were considered spoiled.
points. When estimating the sensory shelf-life of vacuum-packed pork Quality-based methods have been a popular approach for esti-
and beef, Blixt and Borch (2002) asked a trained assessor panel to eval- mating the sensory shelf-life of fresh fruits and vegetables as well.
uate the degree of spoilage odor using a 3-point scale (1 = no off-odor, Artés-Hernández, Rivera-Cabrera, & Kader (2007) estimated the
3 = very strong off-odor). Besides, Fernández-López et al. (2008) and shelf-life of fresh-cut lemons by measuring the overall visual quality
Nunes, Emond, and Brecht (2006) used 5-point structured scales to with a 5-point scale (1: extremely poor, 5: excellent). Liu and Li
evaluate the intensity of sensory attributes for estimating the shelf-life (2006) asked a trained assessor panel to evaluate the color, aroma,
of ostrich steaks and papaya fruit respectively. translucence and general appearance of sliced onions using a
10-point quality scale (1 = excellent fresh, and 10 = extremely de-
3.1.3. Quality rating methods teriorated) and estimated sensory shelf-life as the time at which a
The most popular approach for estimating sensory shelf-life has score of 5 (just acceptable) was reached. Similarly, Medina, Tudela,
been measuring product quality throughout storage. The utilization Marín, Allende, and Gil (2012) used this approach for sensory
of this type of method requires the definition of specifications or qual- shelf-life estimation of minimally processed baby spinach.
ity standards and the selection of criteria to evaluate if products com- Quality-rating methods have also been used for other product catego-
ply with the requirements of the quality standards (Costell, 2002). ries, including dairy and meat products (Conte et al., 2011). However,
A. Giménez et al. / Food Research International 49 (2012) 311–325 317

Table 1 trained assessors (Costa, Conte, Buonocore, Lavorgna, & Del Nobile, 2012;
Scale used by for estimating the sensory shelf-life of fresh cod (Gadus morhua) fillets Mendes et al., 2011; Mohapatra, Bira, Firas, Kerry, & Rodrígues, 2011;
with skin the Quality Index Method (QIM) (Cardenas Bonilla et al., 2007).
Patsias et al., 2006; Rocha & Morais, 2003; Salam, 2007; Siripatrawan &
Quality Description Score Noipha, 2012). Despite the fact that many of these articles only used sen-
parameter sory data as a tool to get information about product stability throughout
Skin Brightness Iridescent pigmentation 0 storage, it is important to advance on the application of good practices
Rather dull 1 in sensory and consumer science in shelf-life studies in order to get accu-
Dull 2
rate information on the sensory changes of food products throughout
Mucus Uniform, thin, transparent 0
Little, thicker, opaque 1 storage and good shelf-life estimations.
Clotted, thick, yellowish 2
Flesh Texture Firm 0 3.2. Acceptability limit methodology
Rather soft 1
Very soft 2
Regulatory agencies do not usually monitor the sensory changes of
Blood Bright red, not present 0
Dull red 1 food products throughout storage. Therefore, the maximum tolerable
Shadowy, brown 2 change in the sensory characteristics of a food product could not be
Odor Fresh, neutral 0 determined based on regulations. In quality-based methods, the fail-
Seaweedy, marine, grass 1
ure criteria for estimating sensory shelf-life is arbitrarily selected by
Sour milk 2
Acetic ammonia 3 the researchers.
Color White, grayish 0 Labuza and Schmidl (1988) reported that the shelf-life of frozen
Some yellowish, a little pinkish 1 foods for consumers is approximately 2 to 3 times longer than
Yellow, over all pink 2 shelf-life estimations based on ‘just noticeable differences’ in sensory
Bright Transparent, bluish 0
characteristics, as determined by a trained assessor panel. On the
Opaque 1
Milky 2 other hand, many authors have estimated the shelf-life of minimally
Gaping No gaping, one longitudinal gaping at the neck part of 0 processed lettuce as the time necessary for the intensity of a sensory
the fillet attribute to reach a score of 50% of the scale (Jacxsens, Devlieghere, &
Slight gaping less than 25% of the fillet 1
Debevere, 2002; Li et al., 2001; Piagentini et al., 2005). However,
Slight gaping, 25–75% of the fillet 2
Deep gaping or slight gaping over 75% of the fillet 3
Lareo et al. (2009) reported that if this criterion was considered to es-
Quality index Sum of all the scores 0–18 timate the sensory shelf-life of butterhead lettuce, only 26% of the
consumers would still purchase the product at the end of its shelf-
life, suggesting that the criterion seemed not strict enough and a
many publications dealing with the application of quality-based methods more conservative criteria would be necessary to assure product
for sensory shelf-life estimation or product stability throughout storage quality and avoid consumer complaints. These examples reflect a
do not follow good practices in sensory evaluation with trained assessor clear distinction between changes on sensory characteristics and con-
panels. sumer perception and show the importance of performing consumer
It is widely recommended that the number of trained assessors for studies in order to establish proper criteria to estimate the sensory
rating tasks should range between 8 and 20 in order to get accurate shelf-life of food products.
and reliable results (Lawless & Heymann, 2010; Stone & Sidel, 2004). According to Hough et al. (2003), food products do not have sen-
However, several studies use a limited number of assessors for estimat- sory shelf lives of their own; rather they depend on the interaction of
ing sensory quality throughout storage. For example, Brash, Charles, the food with the consumer. Therefore, sensory shelf-life is usually
Wright, and Bycroft (1995) asked two observers to measure the overall determined by consumers who find the quality of the product to be
visual quality of asparagus using 9-point quality scale; whereas Mendes, lower than they expected, leading to a rejection to purchase the prod-
Alves Silva, Anacleto, and Cardoso (2011) performed sensory analysis uct again (Labuza & Schmidl, 1988). Using a consumer-based ap-
with 4 experienced assessors on fish quality for estimating the proach, sensory shelf-life of a food product could be regarded as the
shelf-life of octopus; Li, Brackett, Shewfelt, and Beuchat (2001) used length of time during which the product is still accepted by the ulti-
two trained assessors to evaluate the appearance of iceberg lettuce, mate consumer as having the expected level of quality.
and Medina et al. (2012) used a 5 member trained panel for measuring Consumer acceptance could be regarded as an integrated judg-
the visual quality of spinach. ment of the perceived sensory characteristics of a product and its ap-
Another common drawback of many studies dealing with quality- propriateness for an intended use. The most common ways of
rating systems is the lack of appropriate panel training. Although this gathering information about consumers' acceptance of a food product
step is crucial for determining the validity of the results provided by a is asking them to rate their degree of liking using a hedonic scale
trained panel, many studies do not specify the procedures they used to (Lawless & Heymann, 2010). The traditional 9-point hedonic scale de-
train the assessors in rating product quality (e.g. Artés-Hernández et veloped by Peryam and Pilgrim (1957) is the most widely used he-
al., 2007; Li et al., 2001; Liu & Li, 2006; Zhou et al., 2004). Besides, donic scale for investigating consumer liking of food products (Fig. 5).
when panel training is described, many times the number of sessions is The acceptability limit methodology for sensory shelf-life estima-
not enough to get a homogeneous and reliable panel. For example, tion is based on consumer liking data, collected using a hedonic scale.
Siripatrawan and Noipha (2012) only followed one preparatory session Consumers are presented with a set of samples with different stor-
prior to testing odor and appearance attributes of pork sausages, which age times and are asked to score their overall liking using a 9-point
might be insufficient to get accurate results. hedonic scale. In order to estimate sensory shelf-life, a dispersion di-
Despite lack of training could limit the validity of quality data from agram of average overall liking scores against storage time is obtained
trained panels, the most important drawback of many sensory shelf-life and a linear regression is usually performed. Fig. 6 shows the typical
studies is gathering hedonic data with trained assessors instead of objec- linear decrease of overall liking scores as a function of storage time
tive quality ratings. Sensory evaluation textbooks have extensively rec- for hamburger buns. Linear relationships between overall liking
ommend that trained assessors should not measure liking since their scores and storage time have been found for several food products,
hedonic perception is not representative of the perception of a naïve con- including natural passion fruit isotonic drink (De Marchi, Monteiro,
sumer (Lawless & Heymann, 2010; Stone & Sidel, 2004). However, several & Cardello, 2003), apple-baby food (Gámbaro, Ares, et al., 2006);
recent sensory shelf-life studies base their conclusions on liking data from brown pan bread (Giménez et al., 2007), alfajores (Giménez et al.,
318 A. Giménez et al. / Food Research International 49 (2012) 311–325

2008a) and chocolate and carrot cupcakes (Montes Villanueva &


Trindade, 2010), among others.
Sensory shelf-life is determined as the time required for overall
liking scores of the product to fall below a predetermined value. Dif-
ferent criteria have been considered in the literature. Muñoz et al.
(1992) considered an overall liking score of 6.5 as acceptability limit
for quality control specifications. Meanwhile, Gámbaro, Ares, et al.
(2006), Giménez et al. (2007) and Giménez et al. (2008a) used an av-
erage value of 6 (like slightly) in a 9-point hedonic scale as accept-
ability limit considering that is the first score indicating that the
consumer likes the product (c.f. Fig. 5). This approach estimates sen-
sory shelf-life as the time period during which consumers like the
product and therefore they do not just accept to consume the product
but also enjoy it. By applying this criterion to the overall liking scores
of hamburger buns shown in Fig. 6, sensory shelf-life would be esti-
mated in 3.5 days.
Montes Villanueva & Trindade, 2010 considered an average overall
Fig. 6. Consumer average overall liking scores as a function of storage time for ham-
liking score of 5.0 (neither like nor dislike) on a 9-point hedonic scale
burger buns.
as the quality limit. This value is less strict and if applied to the ham-
burger buns data, sensory shelf-life would be estimated in 7 days. In
this example it is clear that the selection of the acceptability limit of consumer overall liking data, considering consumers as blocking fac-
has important implications for the manufacturer since it determines tor and samples as source of variation; and n = number of consumers.
the quality of the product at the end of storage but also influences This methodology assures product quality throughout its whole
commercialization times. shelf-life since it reflects the time when consumers noticed the first sig-
Other less conservative criteria have been considered in different nificant difference in the sensory characteristics of the product with re-
products. For example, when estimating the sensory shelf-life of spect to the fresh one. However, according to Giménez et al. (2007)
trout fillets, Mexis, Chouliara, and Kontominas (2009) considered an estimated shelf lives are sometimes too short and it would be too con-
average score of 4 (dislike slightly) as the lower acceptability limit. servative a criterion to be used by the product manufacturer.
The main problem of estimating sensory-shelf-life considering a A key point that has to be taken into account when designing
predetermined overall liking scored is that several fresh products shelf-life experiments based on acceptability limit methodology is
could have a liking score close or even lower than the acceptability the number of consumers that should evaluate the products. Several
limit. In these cases the criterion proposed by Hough et al. (2002) sensory evaluation textbooks recommend that the number of con-
could be used. These authors determined sensory shelf-life as the sumers necessary to perform a hedonic test ranges from 50 to 300
storage time when the first significant change in overall acceptability (Meilgaard et al., 1991; Stone & Sidel, 2004). Recently, Hough et al.
is detected. At this time, consumers detect the first significant change (2006) provided basic concepts needed to estimate the number of
in the sensory characteristics of the product with respect to the fresh consumers necessary for acceptability studies. In this sense, the
product. The acceptability limit (S) is calculated as the first overall lik- major drawback of some sensory shelf-life experiments is related to
ing score that is significantly different from the overall liking of the the number of consumers. For example, Mastromatteo et al. (2011)
fresh sample using the following equation: and Mastromatteo, Danza, Conte, Muratore, and Del Nobile (2010)
used a panel of 7 consumers to evaluate the overall liking of minimal-
2MSE ly processed kiwifruit and shrimps respectively.
rffiffiffiffiffiffiffiffiffiffiffiffiffi
S ¼ F−Z α ð1Þ
n

where: S = minimum tolerable overall liking score of a stored sample or 3.3. Cut-off point methodology
acceptability limit; F = overall liking score of the fresh sample; Zα =
one-tailed coordinate of the normal curve for α significance level; Despite the importance of consumer data for sensory shelf-life es-
MSE = mean square of the error derived from the analysis of variance timation, to repeatedly perform consumer studies is tedious, time
consuming and expensive (Hough, 2010). In particular, when a sen-
sory shelf-life experiment is performed following a basic storage de-
9 = Like extremely
sign it would be necessary to carry out at least 6 consumer studies
8 = Like very much in different occasions. On the other hand, it is easier to periodically
gather trained assessor panels to evaluate samples throughout a
7 = Like moderately predetermined period of time (Hough et al., 2002) or even to rely
on analytical or instrumental measurements. Whenever possible, it
6 = Like slightly is easier and simpler to estimate the intensity of sensory attributes
using analytical or instrumental measurements. This approach is quite
5 = Neither like nor dislike
simple when working with products which shelf-life is determined by
4 = Dislike slightly texture (Fiszman, Salvador, & Varela, 2005; Gámbaro, Giménez, Ares, &
Gilardi, 2006) or color (Gámbaro, Giménez, et al., 2006; Kong & Chang,
3 = Dislike moderately 2009; Zhou et al., 2004). In the case of flavor or odor attributes it is
more difficult to have an analytical method to monitor their changes
2 = Dislike very much throughout storage. However, some successful approaches have been
reported. For example, flavor changes due to lipid oxidation have been
1 = Dislike extremely
monitored by analytical measurements such as peroxide value or
Fig. 5. Traditional 9-point hedonic scale for determining consumer overall liking of thiobarbituric acid reactive substance (TBARS) (Anacleto et al., 2011;
food products. Gómez & Lorenzo, 2012).
A. Giménez et al. / Food Research International 49 (2012) 311–325 319

Although analytical and instrumental measurements or trained sen- shown in Fig. 7b. The increase in the intensity of sensory attributes has
sory panels are more appropriate for repeated assessments, data would been modeled using zero or first order kinetic models, corresponding
be analytical and not necessarily representative of consumer responses. to linear or exponential models respectively (Hough, 2010).
By correlating data from a consumer panel with those obtained from a This methodology has been used to estimate the sensory shelf-life of
trained panel, the intensity of sensory attributes measured by a trained sunflower oil (Ramírez et al., 2001), powdered milk (Hough et al., 2002),
assessor panel could be used to estimate sensory shelf-life of food prod- dulce de leche (Garitta et al., 2004), apple-baby food (Gámbaro, Ares, et
ucts using failure criteria determined by consumers (Hough, 2010). This al., 2006), human milk replacement formula (Curia & Hough, 2009) and
methodology, developed by Ramírez, Hough, and Contarini (2001) is high hydrostatic pressure processed avocado paste (Jacobo-Velázquez &
called cut-off point, and can be regarded as a combination of intensity Hernández-Brenes, 2011). Besides, failure criteria could be selected
measurement and acceptability limit methodology. This approach en- based on the percentage of consumers rejecting the product instead of
ables working with trained assessors instead of consumers, overcoming considering overall liking scores (Giménez et al., 2007; Lareo et al.,
the limitation of selecting an arbitrary failure criterion. 2009).
Cut-off point methodology requires the identification of the critical
descriptor, i.e. the sensory defect that appears as a result of prolonged
storage and that is responsible for consumer rejection of the product. 3.4. Survival analysis
Critical descriptors could be identified by evaluating samples with differ-
ent storage times (Gámbaro, Giménez, et al., 2006) or by storing samples Shelf-life decisions based on arbitrarily selected acceptability limits
at higher temperature to accelerate sensory deterioration (Garitta, might be taken with caution, as they do not always reflect consumer's
Hough, & Sánchez, 2004; Lareo et al., 2009). Accelerated storage to iden- decision to accept or reject the product. Overall liking scores usually
tify critical descriptors should be taken with care since sensory changes provide little information as to what consumers would normally do
in the product at higher temperatures might not be the same as those when facing the product since the fact that the average overall liking
due to storage time at normal temperatures (Robertson, 2006). This is scores falls below a predetermined value does not mean that, at that
particularly relevant in foods in which phase changes may occur, such time, the consumers would refuse to consume the product (Gámbaro,
as frozen foods or chocolate (Lawless & Heymann, 2010). On the other Giménez, et al., 2006). Giménez et al. (2007) showed that an overall lik-
hand, accelerated storage is a good option for identifying the sensory ing score could imply different proportions of consumer rejecting to
changes which limit the shelf-life of fresh fruits and vegetables or consume the product. These authors reported that an overall liking of
shelf-stable food products. 6 corresponded to 23% of Uruguayan consumers rejecting to consume
Care must be taken when selecting the critical descriptor since brown pan bread, whereas it corresponded to 11% in Spain. Based on
consumers could show different reactions toward similar intensities the results they concluded that Spanish consumers tended to decrease
of different defects. For this reason, it could be a good option to their overall acceptability scores while deciding to accept to consume
keep more than one critical descriptor at this stage of the experiment. the product.
The methodology consists of six basic steps (Hough, 2010): i) In order to estimate sensory shelf-life based on consumer rejec-
preparation of a set of samples with increasing intensity of the critical tion of a food product, survival analysis could be applied. Survival
sensory defect, ii) evaluation of the intensity of the sensory defect of analysis is set of statistical procedures applicable for the analysis of
the set of samples by a trained assessor panel, iii) determination of time until an event of interest occurs; being extensively used in clin-
consumer liking of the set of samples, iv) regression of overall liking ical studies, epidemiology, biology, sociology, and reliability studies
as a function of the intensity of the sensory defect, v) regression of (Klein & Moeschberger, 1997; Meeker & Escobar, 1998). When ap-
the intensity of the sensory defect as a function of storage time, and plied to sensory shelf-life estimations, this methodology focuses the
vi) estimation of the sensory-shelf-life. shelf-life risk on consumers' rejection of the product rather than on
Different approaches could be used to get a set of samples with in- the product deteriorating (Hough et al., 2003).
creasing intensity of a sensory defect, including using different batches
of products with different storage times (Gámbaro, Giménez, et al., Table 2
2006), storing samples at higher temperatures (Ramírez et al., 2001) Examples of different approaches used for obtaining samples with different intensities
or modifying fresh samples by adding different concentrations of a ref- of a defect for sensory shelf-life estimation using the cut-off point methodology.
erence compound (Giménez, Ares, & Gámbaro, 2008b; Hough et al., Product Sensory Approach Reference
2002). Table 2 summarizes different approaches used for the generation defect
of sensory defects in different products. After samples with different in-
Sunflower oil OxidizedStoring samples at 60 °C for Ramírez et al.
tensities of the sensory defect have been generated, consumers are flavor different times (2001)
presented with the sample set and are asked to score their overall liking Apple-baby Color Different industrial batches Gámbaro, Ares, et
using a 9-point hedonic scale. Additionally, the trained assessor panel is food stored for 0, 9, 26 and 36 months al. (2006)
asked to rate the intensity of the sensory defect using unstructured in- at 25 °C
Reconstituted Acid Addition of lactic acid to a fresh Hough et al. (2002)
tensity scales. In order to estimate the cut-off point, a dispersion dia- milk sample
gram of overall liking scores against the intensity of the defect is powder Caramel Addition of caramel flavoring to a
obtained and a regression performed. Linear regressions have been fresh sample
commonly reported (Garitta et al., 2004; Hough et al., 2002; Ramírez Dulce de Burnt Addition of sugar heated till first Garitta et al.
leche smoke to a fresh sample (2004)
et al., 2001). Fig. 7a shows a typical plot of overall liking scores as a func-
Plastic Mixing of a fresh sample with
tion of the intensity of a sensory defect for orange juice. Using the linear one heated in a polystyrene pot
regression, the cut-off point is determined as the intensity of the defect during 24 h at 80 °C
that corresponds to a predetermined overall liking score, as in the Dulce de Sandiness Mixing fresh dulce de leche with Giménez et al.
Acceptability limit methodology (c.f. Section 3.2). The most common leche lactose crystals of different sizes (2008b)
Fluid human Metallic Addition of a ferrous sulfate Curia and Hough
approach is to consider the first significant difference in overall liking milk solution to fresh fluid human (2009)
as minimum acceptability. Considering this criterion in the example milk
depicted in Fig. 7a, for a minimum acceptability of 5.8, the cut-off Avocado Sour Addition of citric acid to freshly Jacobo-Velázquez
point for off-flavor intensity in orange juice corresponds to 2.4. This paste processed avocado paste and
Rancid Addition of hexanal to freshly Hernández-Brenes
cut-off point should be used to estimate the sensory shelf-life of orange
flavor processed avocado paste (2011)
juice as the time necessary for off-flavor to reach an intensity of 2.4, as
320 A. Giménez et al. / Food Research International 49 (2012) 311–325

exactly, giving as a result censored data (Hough et al., 2003). If con-


sumers are presented with six samples with different storage times,
three types of censoring could be identified. If the consumer rejects
the sample at the first storage time considered, the shelf-life for that con-
sumer (T) is not observed since it is shorter than the first storage time
(T≤t1) and the data is left censored (e.g. consumer 50 in Table 3). If a
consumer accepts to consume the sample stored for t2 and rejects the
sample stored for t3 (e.g. consumer 1 in Table 3) the exact time at
which he/she rejects the product (i.e. shelf-life) occurs between t2 and
t3 (t2b T≤t3) and the data is interval censored. Finally, if a consumer ac-
cepts all samples, then rejection is not observed (e.g. consumer 2 in
Table 3), shelf-life is longer than the last storage time considered
(T>t6) and the data is right censored. Sensory shelf-life is estimated con-
sidering acceptance/rejection data of each individual consumer.
Defining a random variable T as the storage time at which a con-
sumer rejects the sample, the survival function S (t) can be defined
as the probability of a consume accepting a product stored for a
time period longer than t, that is S (t) = P (T > t). Alternatively, the
cumulative distribution function F (t) can be defined as the probabil-
ity of a consumer rejecting a product stored for a time period shorter
than t, that is F (t) = P (T ≤ t). F (t) can be interpreted as the propor-
tion of consumers who will reject a food product stored for a period of
time shorter than t (Hough et al., 2003).
A non-parametric estimation of the survival functional could be
obtained though the likelihood function, which is a mathematical ex-
pression that describes the joint probability of obtaining the given ob-
servations for the n consumers (Klein & Moeschberger, 1997):

L ¼ ∏ Sðr i Þ⋅ ∏ ð1−SðIi ÞÞ⋅ ∏ ðSðIi Þ−Sðr i ÞÞ ð2Þ


i∈R i∈L i∈I

where R is the set of right-censored observations, L the set of


left-censored observations, and I is the set of interval-censored
observations.
Alternatively, parametric models could be used to obtain precise
Fig. 7. a) Regression of consumer average overall liking as a function of off-flavor
intensity in orange juice and calculation of the cut-off point. b) Sensory shelf-life esti- estimates of the survival function by assuming Weibull or log-linear
mation based on the cut-off point for off-flavor intensity in orange juice. distributions for the survival times (Klein & Moeschberger, 1997).
If the log-normal distribution is chosen for T, the survival function
is given by:
When applying survival analysis, consumers are asked to try a set
of samples with different storage times and answer “yes” or “no” to 
lnðt Þ−μ

the question “Would you normally consume this product?”. It is usu- Sðt Þ ¼ 1−ϕ ð3Þ
σ
ally explained to consumers that they have to indicate if they would
consume the product after buying it or if it was served to them at where ϕ (•) is the standard normal cumulative distribution function,
their homes (Hough et al., 2003). and μ and σ are the model's parameters.
The number of consumers considered in most sensory shelf-life ex- Meanwhile, if the Weibull distribution is chosen, the survival func-
periments dealing with survival analysis has been close to 50 (Giménez tion is given by:
et al., 2007; Hough et al., 2003; Varela et al., 2005). Hough, Calle, Serrat,
and Curia (2007) provided recommendations for the number of con- lnðt Þ−μ
 
sumers necessary for making sensory shelf-life estimations based on Sðt Þ ¼ Sev ð4Þ
σ
survival analysis statistics. These authors concluded that in many occa-
sions the recommended number of consumers would be close to 120, where Ssev (•) is the survival function of the smallest extreme value
higher to that usually considered. distribution: Ssev (w) = exp (− e w), and μ and σ are the model's
It is important to take into account that in this approach it is neces- parameters.
sary that each consumer try all the samples with different storage times.
Thus, consumers could evaluate the whole sample set from a reversed
storage design in a single session (Giménez et al., 2007; Hough et al., Table 3
2003; Østli, Esaiassen, Garitta, Nøstvold, & Hough, in press) or they Example of the data matrix used for analyzing data from survival analysis and type of
censoring.
could evaluate samples in different sessions, according to a basic storage
design (Ares, Parentelli, Gámbaro, Lareo, & Lema, 2006; Lareo et al., Consumer Storage time Censoring
2009). t1 t2 t3 t4 t5 t6
Table 3 exemplifies how data is prepared for analysis. For each
1 Yes Yes No No No No Interval
consumer the acceptance/rejection data is included in a row that in- 2 Yes Yes Yes Yes Yes Yes Right
dicates if he/she accepted (yes) or rejected (no) the sample from … … … … … … … …
each storage time. Due to the fact that consumers evaluate a limited n No No No No No No Left
number of samples with different storage times, the exact storage Yes indicates that the consumer accepted to consume the sample, whereas No
time at which he/she rejects the product could not be observed indicates that the consumer rejected it.
A. Giménez et al. / Food Research International 49 (2012) 311–325 321

The parameters of the log-linear or Weibull model are obtained by (Manzocco & Lagazio, 2009), avocado and mango pulps (Jacobo-
maximizing the likelihood function (Eq. (2)), substituting S (t) in Eq. (2) Velázquez et al., 2010) and fresh cod fillets (Østli et al., in press).
by the expressions given in Eqs. (3) or (4), respectively (Hough et al., Ares, Giménez, and Gámbaro (2008) evaluated the influence of the
2003). Once the likelihood function is obtained for a given model, special- stage of the decision-making process (purchase or consumption stage)
ized software can be used to estimate the parameters (μ and σ) that max- considered for estimating the sensory shelf-life of minimally processed
imize the likelihood function for the given experimental data set. lettuce, based on the fact that consumer evaluative criteria may change
In order to select the best distribution for the survival function, the (Gardial, Clemons, Woodruff, Schumann, & Burns, 1994). Ares et al.
different parametric models are assayed and visual assessment of how (2008) asked a consumer panel to evaluate the appearance of minimally
they adjust to the non-parametric estimation is performed (Hough et processed lettuce and to answer “yes” or “no” to the questions: “Imagine
al., 2003). An example of how standard distributions adjust to experi- you are in a supermarket, you want to buy a minimally processed lettuce,
mental data is shown in Fig. 8. Log-normal and Weibull distributions and you find a package of lettuce with leaves like this, would you nor-
are the most common options in sensory shelf-life estimations (Ares mally buy it?” and “Imagine you have this leaf of lettuce stored in your
et al., 2006; Gámbaro, Giménez, et al., 2006; Giménez et al., 2007; refrigerator, would you normally consume it?”. Shelf lives estimated con-
Hough et al., 2003). In the example shown in Fig. 8, log‐normal distribu- sidering rejection to purchase were significantly lower than those esti-
tion showed the best fit. mated considering rejection to consume. These results indicate that
After the distribution of the survival function is chosen, the maxi- consumers are harsher when selecting a product at purchase stage
mum likelihood estimates of the parameters are calculated and used than at consumption stage. The effect was attributed to the fact that
to graph the proportion of consumers rejecting the product as a func- when they considered purchase stage they were thinking about storing
tion of storage time, as shown in Fig. 9. In order to estimate shelf-life, the product before consuming it, or that when they are at their homes
the probability of a consumer rejecting a product (F (t)) must be chosen. they are more tolerant to defects because they have already bought the
Gacula and Singh (1984) mentioned a nominal shelf-life value consider- product and they do not want to throw it away. Therefore, Ares et al.
ing 50% rejection. Cardelli and Labuza (2001) used this criterion in cal- (2008) concluded that in order to assure product quality, shelf-life of
culating shelf-life of coffee, whereas Hough et al. (2003) recommended minimally lettuce should be estimated considering consumers' rejection
this percentage when estimating the sensory shelf-life of yogurt. Ares et to purchase instead of rejection to consume, as traditionally has been
al. (2006), Gámbaro, Ares, et al. (2006), Giménez et al. (2007) used 25% done. This seems particularly important when sensory shelf-life is limit-
rejection to estimate the shelf-life of baked products. This means that if ed by appearance changes that can be seen by consumers when buying
a consumer tries the product at the end of its shelf-life, there is a 25% the product. This approach has been used by Østli et al. (in press) to es-
probability that he will reject it. Considering that few consumers will timate the shelf-life of fresh cod fillets.
taste the product near the end of its shelf-life, and that of the few that Giménez et al. (2008a) proposed another modification of the usual
do 75% will still find the product acceptable, this value of F (t) = 25% methodology, focusing the risk on consumers' disliking the product
seems reasonable from a practical point of view. By applying this crite- instead of on consumers' rejecting it. These authors estimated senso-
rion to the data shown in Fig. 9, the shelf-life of hamburger buns would ry shelf-life of pan bread and alfajores applying survival analysis sta-
be estimated in 3 days. tistics on overall liking scores by performing a data transformation. If
Survival analysis methodology has been used to estimate the senso- the acceptability score of a consumer for a sample was 1–5, then the
ry shelf-life of different food products, including yogurt (Hough et al., rating was transformed to the word “no”, indicating that the consum-
2003), Fuji apples (Varela et al., 2005), apple-baby food (Gámbaro, er disliked the product. In contrast, if a consumer's score for a sample
Ares, et al., 2006), minced beef (Hough, Garitta, & Gómez, 2006), shiita- was 6–9, it was replaced by the word “yes”, indicating that the con-
ke mushrooms (Ares et al., 2006), brown pan bread (Giménez et al., sumer liked the product. Shelf lives estimated considering 50% of
2007), pears (Salvador et al., 2007), muffins (Baixauli, Salvador, & the consumers disliking the product were shorter than those estimated
Fiszman, 2008), butterhead lettuce (Lareo et al., 2009), coffee brew considering consumers' rejection to consume, which was attributed to

Fig. 8. Comparison of 6 distribution models for the survival function for estimating the probability of consumer rejecting samples with different storage times.
322 A. Giménez et al. / Food Research International 49 (2012) 311–325

where R is the set of right-censored observations and L the set of


left-censored observations.
In their study, Araneda et al. (2008) used a Weibull distribution
for modeling the rejection function and estimated the sensory
shelf-life of ready-to-eat lettuce as 11 and 15 days for 25% and 50%
consumer rejection percentage, respectively.
Based on simulation studies Libertino et al. (2011) concluded that,
for a given number of parameters, a total of 300 consumers are neces-
sary to obtain valid shelf-life estimations from survival analysis calcu-
lations where each consumer evaluates a single sample. Considering a
total of 6 storage times, it would be necessary that 50 consumers try
the samples at each storage time.

4. Methodological recommendations

The most popular approach for estimating the sensory shelf-life of


Fig. 9. Parametric estimation of the percentage of consumers rejecting the product as a
food products continues to be the evaluation of product quality
function of storage time, assuming a log‐normal distribution for the survival function, throughout storage with a trained assessor panel. The main disadvan-
for hamburger bun samples. tage of this approach is that failure criteria are arbitrarily selected, as
extensively discussed by Hough and Garitta (2012) in their review.
the fact that a proportion of consumers might dislike the sample but still Therefore, shelf-life is determined as the time necessary for reaching
answer “yes” when asked if they would consume the product at their a predetermined change in the sensory characteristics of the product,
homes. Based on these results the authors concluded that estimating but consumer perception of the product when it reaches the end of its
sensory shelf-life based on the percentage of consumers disliking the shelf-life is unknown.
product consisted of a conservative criterion to assure product quality This approach could lead to inaccurate sensory shelf-life estima-
throughout its storage. tions. For example, if difference testing is used, sensory shelf-life esti-
mations only tell the moment at which the product changes its
sensory characteristics. However, sensory differences detected by the
3.4.1. Current status survival analysis
trained panel are usually small and might have little relevance to the
When estimating sensory shelf-life using survival analysis, each
sensory quality as perceived by consumers (Griffiths, 1985). On the
consumer should evaluate all the samples with different storage times
other hand, sensory shelf-life estimations based on arbitrarily selected
(usually 6 or 7) in a single session. This could be easily achieved follow-
attribute intensities or product quality could lead to high consumer dis-
ing a reversed storage design (Hough, 2010). However, in many occa-
satisfaction at the end of the shelf-life period, as suggested by Lareo et
sions it is not feasible to use this design, being necessary to use a basic
al. (2009). Consequently, inaccurate shelf-life estimations could cause
storage design and therefore the same consumers should perform re-
a decrease in consumers' confidence in the brand and in the store that
peated tests for each of the storage times considered (Ares et al.,
sells it, and rejection of that particular product in following purchase oc-
2006; Lareo et al., 2009). As discussed in Section 2.1, this type of ap-
casions, which could cause important economic losses for manufactur-
proach is time consuming, expensive and could lead to biased results.
ing companies.
Moreover, it would be more representative of a real-life situation for
For all the above-mentioned reasons, consumer studies are the best
consumers to try a single sample and to indicate if they would consume
alternative for estimating the sensory shelf-life of food products. Among
it or not (Libertino, López Osornio, & Hough, 2011).
the different alternatives for estimating sensory shelf-life based on con-
In this context, Araneda, Hough, and Wittig de Penna (2008) de-
sumer data, survival analysis is clearly the recommended approach
veloped current-status survival analysis, a method for estimating sen-
since it better reflects what consumers will have to decide when facing
sory shelf-life using survival analysis for situations in which each
a product at their homes: they will accept or reject to consume it. De-
consumer evaluates only one sample corresponding to one storage
spite the fact that it has been recently developed (Hough et al., 2003),
time current-status data.
a large number of studies have proven its applicability in a wide range
Araneda et al. (2008) applied current-status survival analysis to es-
of food products. When it is possible to use reversed storage design sur-
timate the sensory shelf-life of ready-to-eat lettuce. The authors
vival analysis is recommended, whereas for products that require the
performed 6 studies with 50–52 consumers each, in which they asked
use of basic storage design, current-status data seems a promising
them to evaluate the appearance, texture and flavor of the lettuce and
to answer “yes” or “no” to the question “Would you regularly consume
this lettuce?”. In this approach, survival analysis only right and left cen- Table 4
soring exists since each consumer evaluates only one product which has Example of the data matrix used for analyzing data from current status survival analy-
sis and type of censoring.
been stored for the period of time corresponding to t. If a consumer
evaluates a sample stored for t2 and rejects it (e.g. consumer 52 in Consumer Storage time Response Censoring
Table 4), his/her data is left-censored since the exact rejection time 1 t1 Yes Right
(T) is shorter than t3 (T≤ t3). On the other hand, if a consumer tries a 2 t1 Yes Right
sample stored for t2 and accepts it (e.g. consumer 51 in Table 4), his/ … … … …
50 t1 No Left
her data is right-censored since the exact rejection time is longer than
51 t2 Yes Right
t2 (T>t2). Data analysis is performed as in survival analysis on the 52 t2 No Left
data matrix shown in Table 4. The models used are similar to those de- … … … …
scribed in Section 3.4 with the exception that the likelihood function 100 t2 No Left
should be replaced by the following expression: 101 t3 No Left
… … … …
n t6 No Left

L ¼ ∏ Sðr i Þ⋅ ∏ ð1−SðIi ÞÞ ð5Þ Yes indicates that the consumer accepted to consume the sample, whereas No
i∈R i∈L indicates that the consumer rejected it.
A. Giménez et al. / Food Research International 49 (2012) 311–325 323

alternative. Considering that only two applications of current-status would not buy the product again. This area of research is a challenging
survival analysis have been reported, further research dealing with for both the academy and food industries.
basic aspects of the methodology is needed. Another relevant issue that has not been addressed in sensory
Another interesting and cost-effective alternative for sensory shelf-life shelf-life estimation is the influence of non-sensory variables. According
estimation is the application of cut-off methodology, which requires a sin- to Rozin and Tuorila (1993) food acceptance cannot be understood with-
gle consumer study for the definition of failure criteria for attribute inten- out consideration of context; being regarded as the set of events and ex-
sity. This approach is particularly useful when evaluating the influence of periences that are not part of the reference event but have some
formulation variables on sensory shelf-life or when repeatedly assessing relationship to it. Several contextual variables such as convenience,
the shelf-life of different industrial batches. Moreover, this methodology price, branding, time of day, eating situation, and advertising informa-
is particularly useful for the identification of the sensory attributes that tion, have been reported to have a large influence on consumer percep-
limit the shelf-life of food products. An example of the application of tion (Jaeger, 2006). However, traditionally, when evaluating samples
this approach is reported by Jacobo-Velázquez and Hernández-Brenes with different storage times, contextual factors have been controlled
(2011). These authors determined the cut-off point of sour and rancid and consumers have tried samples blind, so that differences between
flavors in avocado paste and used a stepwise logistic regression for deter- products are only due to their sensory characteristics. However, if con-
mining their contribution to consumer rejection. sumers were informed of non-sensory characteristics of the products
Furthermore, it is important to highlight that sensory shelf-life is de- their perception could largely differ, leading to changes in estimated
pendent on the methodology selected for the estimation due to the fact shelf-life dates. Østli et al. (in press) reported that capture-date informa-
that they focus on different aspects of product deterioration. Methodol- tion caused a significant decrease in the shelf-life of fresh cod fillets,
ogies based on the evaluation of changes in the product sensory charac- which indicates that prior beliefs about product freshness could override
teristics are product-focused and provide shelf-life estimations that the influence of sensory changes in determining consumer rejection to
reflect a predetermined sensory change in the stored product with re- buy a fresh product. On the other hand, consumers might be more toler-
spect to the fresh one. On the other hand, consumer-based methodolo- ant when finding a defective product of their usual brand than when try-
gies rely on consumer reaction to the sensory changes that occur as a ing a product of an unknown brand, which suggests that unknown
result of prolonged storage. In these methodologies sensory shelf-life brands or products should be more strict when estimating the sensory
is estimated as the time necessary for consumers to decrease their liking shelf-lives of their products because many consumers do not know the
or acceptance to a certain degree. However, different consumer-based product and therefore would not re-buy it unless they really like it. For
methodologies might yield different sensory shelf-life estimations since this reason, studies addressing the influence of non-sensory variables
they rely on different aspects of consumer perception. Hedonic scales such as brand or price on sensory-shelf-life estimations could provide
could be used to estimate sensory shelf-life based on a change on con- valuable information to the industry.
sumer hedonic perception of the products. However, these approaches
do not always reflect consumer behavior when deciding whether to ac-
cept or reject a certain product for its consumption or purchase. For ex- Acknowledgments
ample, Giménez et al. (2007) reported that Spanish consumers showed
a tendency to decrease their overall liking scores for brown pan bread The authors are indebted to CSIC (Comisión Sectorial de Investigación
while accepting to consume the product. Científica, Universidad de la República) for the financial support provided
to a research project dealing with sensory shelf-life estimation, and to
Comisión Académica de Posgrado (Universidad de la República) for the
5. Challenges and suggestions for further research PhD scholarship granted to Ana Giménez.

The main question that arises when estimating sensory shelf-life is


how to select the failure criteria or cut-off point. During this review it References
has been extensively argued that the best approach is to select failure
Alkadamany, E., Toufeili, Khattar, M., Abou-Jawdeh, Y., Harakeh, S., & Haddad, T.
criteria based on consumer perception. In this sense, sensory shelf-life (2002). Determination of shelf life of concentrated yogurt (labneh) produced by
has been estimated as the time at which overall liking reaches a in-bag straining of set yogurt using hazard analysis. Journal of Dairy Science, 85,
1023–1030.
predetermined value (Gámbaro, Ares, et al., 2006; Giménez et al.,
Anacleto, P., Teixeira, B., Marques, P., Pedro, S., Nunes, M. L., & Marques, A. (2011).
2007) or, when working with survival analysis, as the time at which Shelf-life of cooked edible crab (Cancer pagurus) stored under refrigerated condi-
25 or 50% of the consumers reject to consume the product (Ares et al., tions. LWT — Food Science and Technology, 44, 1376–1382.
2006; Giménez et al., 2007; Hough et al., 2003). However, it is not Araneda, M., Hough, G., & Wittig de Penna, E. (2008). Current-status survival analysis
methodology applied to estimating sensory shelf life of ready-to-eat lettuce
clear what the implications of selecting different failure criteria for (Lactuca sativa). Journal of Sensory Studies, 23, 162–170.
different product categories are. If consumers found a deteriorated Ares, G., Giménez, A., & Gámbaro, A. (2008). Sensory shelf life estimation of minimally
product within its labeled shelf-life date, they could simply discard processed lettuce considering two stages of consumers' decision-making process.
Appetite, 50, 529–535.
the product without any further action or reject to purchase that partic- Ares, G., Parentelli, C., Gámbaro, A., Lareo, C., & Lema, P. (2006). Sensory shelf life of
ular product again, causing an economic loss for the manufacturer. A shiitake mushrooms stored under passive modified atmosphere. Postharvest Biolo-
small proportion of the consumers might decide to phone up the man- gy and Technology, 41, 191–197.
Artés-Hernández, F., Rivera-Cabrera, F., & Kader, A. A. (2007). Quality retention and po-
ufacturer to complain. According to Saguy and Peleg (2009) a common tential shelf life of fresh-cut lemons as affected by cut type and temperature.
assumption is that the ratio of complainers and non-complainers when Postharvest Biology and Technology, 43, 245–254.
finding a deteriorated product is 1 to 65–100. Therefore, information Baixauli, R., Salvador, A., & Fiszman, S. M. (2008). Textural and colour changes during
storage and sensory shelf life of muffins containing resistant starch. European
relating failure criteria, the number of consumers rejecting to buy the
Food Research and Technology, 226, 523–530.
product again and the number of complaints received by the food com- Bishop, J. R., & White, C. H. (1986). Assessment of dairy product quality and potential
pany could contribute to developing good practices for sensory shelf-life — A review. Journal of Food Protection, 49, 739–753.
Blixt, Y., & Borch, E. (2002). Comparison of shelf life of vacuum-packed pork and beef.
shelf-life estimation. Peleg, Normand, and Corradini (2011) described
Meat Science, 60, 371–378.
a method based on a ‘Fermi solution’ for relating probability of receiving Bono, G., & Badalucco, C. (2012). Combining ozone and modified atmosphere packag-
consumer complaints and the number of spoiled product units available ing (MAP) to maximize shelf-life and quality of striped red mullet (Mullus
in the market, which could be potentially applied for estimating the rela- surmuletus). LWT — Food Science and Technology, 47, 500–504.
Brash, D. W., Charles, C. M., Wright, S., & Bycroft, B. L. (1995). Shelf-life of stored aspar-
tionship between different failure criteria and the number of complaints agus is strongly related to postharvest respiratory activity. Postharvest Biology and
received by the manufacturer or even the number of consumers that Technology, 5, 77–81.
324 A. Giménez et al. / Food Research International 49 (2012) 311–325

Cardelli, C., & Labuza, T. P. (2001). Application of Weibull hazard analysis to the deter- Hyde, K., Smith, A., Smith, M., & Henningsson, S. (2001). The challenge of waste
mination of the shelf life of roasted and ground coffee. Lebensmittel-Wissenschaft & minimisation in the food and drink industry: A demonstration project in East An-
Technologie, 34, 273–278. glia, UK. Journal of Cleaner Production, 9, 57–64.
Cardello, A. V. (1995). Food quality: Relativity, context, and consumer expectations. IFST (1993). Shelf life of foods: Guidelines for its determination and prediction. London:
Food Quality and Preference, 6, 163–170. Institute of Food Science and Technology.
Cardenas Bonilla, A. C., Sveinsdottir, K., & Martinsdottir, E. (2007). Development of Jacobo-Velázquez, D. A., & Hernández-Brenes, C. (2011). Sensory shelf-life limiting fac-
Quality Index Method (QIM) scheme for fresh cod (Gadus morhua) fillets and ap- tor of high hydrostatic pressure processed avocado paste. Journal of Food Science,
plication in shelf life study. Food Control, 18, 352–358. 76, S388–S395.
Conte, A., Brescia, I., & Del Nobile, M. A. (2011). Lysozyme/EDTA disodium salt and Jacobo-Velázquez, D. A., Ramos-Parra, P. A., & Hernández-Brenes, C. (2010). Survival
modified-atmosphere packaging to prolong the shelf life of Burrata cheese. Journal analysis applied to the sensory shelf-life dating of high hydrostatic pressure
of Dairy Science, 94, 5289–5297. processed avocado and mango pulps. Journal of Food Science, 75, S286–S291.
Costa, C., Conte, A., Buonocore, G. G., Lavorgna, M., & Del Nobile, M. A. (2012). Jacxsens, L., Devlieghere, F., & Debevere, J. (2002). Temperature dependence of
Calcium-alginate coating loaded with silver–montmorillonite nanoparticles to shelf-life as affected by microbial proliferation and sensory quality of equilibri-
prolong the shelf-life of fresh-cut carrots. Food Research International, 48, um modified atmosphere packaged fresh produce. Postharvest Biology and Tech-
164–169. nology, 26, 59–73.
Costell, E. (2002). A comparison of sensory methods in quality control. Food Quality and Jaeger, S. R. (2006). Non-sensory factors in sensory science research. Food Quality and
Preference, 13, 34–353. Preference, 17, 132–144.
Cruz, A. G., Walter, E. H. M., Silva Cadena, R., Faria, J. A. F., Bolini, H. M. A., Pinheiro, H. P., Jeevanandam, K., Kakatkar, A., Doke, S. N., Bongirwar, D. R., & Venugopal, V. (2001). In-
et al. (2010). Survival analysis methodology to predict the shelf-life of probiotic fluence of salting and gamma irradiation on the shelf-life extension of threadfin
flavored yogurt. Food Research International, 43, 1444–1448. bream in ice. Food Research International, 34, 739–746.
Curia, A. V., & Hough, G. (2009). Selection of a sensory marker to predict the sensory shelf Kilcast, D. (2000). Sensory evaluation methods for shelf-life assessment. In D. Kilcast, &
life of a fluid human milk replacement formula. Journal of Food Quality, 32, 793–809. P. Subramaniam (Eds.), The stability and shelf-life of food (pp. 79–105). Boca Raton,
De Marchi, R., Monteiro, M., & Cardello, M. H. A. B. (2003). Avaliação da vida-de-prateleira LF: CRC/Woodhead.
de um isotônico natural de maracujá (Passiflora edulis Sims f. flavicarpa Deg.). Brazil- Klein, J. P., & Moeschberger, M. L. (1997). Survival analysis, a self learning text. New
ian Journal of Food Technology, 6, 291–300. York: Springer-Verlag.
Dethmers, A. E. (1979). Utilizing sensory evaluation to determine product shelf life. Kong, F., & Chang, S. K. C. (2009). Statistical and kinetic studies of the changes in soy-
Food Technology, 33(9), 40–43. bean quality during storage as related to soymilk and tofu making. Journal of
Fernández-López, J., Sayas-Barberá, E., Muñoz, T., Sendra, E., Navarro, C., & Food Science, 74, S81–S89.
Pérez-Álvarez, J. A. (2008). Effect of packaging conditions on shelf-life of ostrich Labuza, T. P., & Schmidl, M. K. (1988). Use of sensory data in the shelf life testing of foods:
steaks. Meat Science, 78, 143–152. Principles and graphical methods for evaluation. Cereal Foods World, 33, 193–206.
Fiszman, S. M., Salvador, A., & Varela, P. (2005). Methodological developments in bread Labuza, T. P., & Szybist, L. M. (2001). Open dating of foods. Trumbull, CT: Food & Nutri-
staling assessment: Application to enzyme-supplemented brown bread. European tion Press.
Food Research and Technology, 221, 616–623. Lareo, C., Ares, G., Ferrando, L., Lema, P., Gámbaro, A., & Soubes, M. (2009). Influence of
Freitas, M. A., & Costa, J. C. (2006). Shelf life determination using sensory evaluation temperature on shelf life of butterhead lettuce leaves under passive modified at-
scores: A general Weibull modeling approach. Computers & Industrial Engineering, mosphere packaging. Journal of Food Quality, 32, 240–261.
51, 652–670. Lawless, H. T., & Heymann, H. (2010). Sensory evaluation of food. Principles and practices
Gacula, M. C. (1975). The design of experiments for shelf life study. Journal of Food Sci- (2nd ed.). New York: Springer.
ence, 40, 399–403. Li, Y., Brackett, R. E., Shewfelt, R. L., & Beuchat, L. R. (2001). Changes in appearance and
Gacula, M. C., & Kubala, J. J. (1975). Statistical models for shelf life failures. Journal of natural microflora on iceberg lettuce treated in warm, chlorinated water and then
Food Science, 40, 404–409. stored at refrigeration temperature. Food Microbiology, 18, 299–308.
Gacula, M. C., Jr., & Singh, J. (1984). Statistical methods in food and consumers research. Libertino, L. M., López Osornio, M. M., & Hough, G. (2011). Number of consumers nec-
New York: Academic Press. essary for survival analysis estimations based on each consumer evaluating a sin-
Gámbaro, A., Ares, G., & Giménez, A. (2006). Shelf-life estimation of apple-baby food. gle sample. Food Quality and Preference, 22, 24–30.
Journal of Sensory Studies, 21, 101–111. Liu, F., & Li, Y. (2006). Storage characteristics and relationships between microbial
Gámbaro, A., Giménez, A., Ares, G., & Gilardi, V. (2006). Influence of enzymes on the growth parameters and shelf life of MAP sliced onions. Postharvest Biology and
texture of brown pan bread. Journal of Texture Studies, 37, 300–314. Technology, 40, 262–268.
Gardial, S. F., Clemons, D. S., Woodruff, R. B., Schumann, D. W., & Burns, M. J. (1994). Manzocco, L., & Lagazio, C. (2009). Coffee brew shelf life modelling by integration of ac-
Comparing consumer's recall of prepurchase and postpurchase product evaluation ceptability and quality data. Food Quality and Preference, 20, 24–29.
experiences. Journal of Consumer Research, 20, 548–560. Martínez, I., Ares, G., & Lema, P. (2008). Influence of cut and packaging film on sensory
Garitta, L., Hough, G., & Sánchez, R. (2004). Sensory shelf life of dulce de leche. Journal quality of fresh-cut butterhead lettuce (Lactuca sativa L., cv. Wang). Journal of Food
of Dairy Science, 87, 1601–1607. Quality, 31, 48–66.
Giménez, A., Ares, G., & Gámbaro, A. (2008a). Survival analysis to estimate sensory Martínez, C., Mucci, A., Santa Cruz, M. J., Hough, G., & Sánchez, R. (1998). Influence of
shelf life using acceptability scores. Journal of Sensory Studies, 23, 571–582. temperature, fat content and package material on the sensory shelf-life of com-
Giménez, A., Ares, G., & Gámbaro, A. (2008b). Consumer perception of sandiness in mercial mayonnaise. Journal of Sensory Studies, 13, 331–346.
dulce de leche. Journal of Sensory Studies, 23, 171–185. Mastromatteo, M., Conte, A., & Del Nobile, M. A. (2011). Combined effect of active coat-
Giménez, A., Varela, P., Salvador, A., Ares, G., Fiszman, S., & Garitta, L. (2007). Shelf life ing and MAP to prolong the shelf life of minimally processed kiwifruit (Actinidia
estimation of brown pan bread: A consumer approach. Food Quality and Preference, deliciosa cv. Hayward). Food Research International, 44, 1224–1230.
18, 196–204. Mastromatteo, M., Danza, A., Conte, A., Muratore, G., & Del Nobile, M. A. (2010).
Gómez, M., & Lorenzo, J. M. (2012). Effect of packaging conditions on shelf-life of fresh Shelf life of ready to use peeled shrimps as affected by thymol essential oil
foal meat. Meat Science, 91, 513–520. and modified atmosphere packaging. International Journal of Food Microbiology,
Griffiths, N. (1985). Sensory analysis in measuring shelf-life. Food, Flavours, Ingredients, 144, 250–256.
Packaging and Processing, 7(9), 47–48. Medina, M. S., Tudela, J. A., Marín, A., Allende, A., & Gil, M. I. (2012). Short postharvest
Harcar, T., & Karakaya, F. (2005). A cross-cultural exploration of attitudes toward prod- storage under low relative humidity improves quality and shelf life of minimally
uct expiration dates. Psychology & Marketing, 22, 353–371. processed baby spinach (Spinacia oleracea L.). Postharvest Biology and Technology,
Heldman, D. R., & Hartel, R. W. (1997). Principles of food processing. Gaithersburg, 67, 1–9.
Maryland: Aspen Publishers, Inc. Meeker, W. Q., & Escobar, L. A. (1998). Statistical methods for reliability data. New York:
Hough, G. (2010). Sensory shelf life estimation of food products. Boca Raton, FL: CRC John Wiley & Sons.
Press. Meilgaard, M., Civille, G. V., & Carr, B. T. (1991). Sensory evaluation techniques. Boca
Hough, G., Calle, M. L., Serrat, C., & Curia, A. (2007). Number of consumers necessary for Raton, Florida: CRC Press, Inc.
shelf life estimations based on survival analysis statistics. Food Quality and Prefer- Mena, C., Adenso-Diaz, B., & Yurt, O. (2011). The causes of food waste in the supplier–
ence, 18, 771–775. retailer interface: Evidences from the UK and Spain. Resources, Conservation and
Hough, G., & Garitta, L. (2012). Methodology for sensory shelf-life determination: A re- Recycling, 55, 648–658.
view. Journal of Sensory Studies, 27, 137–147. Mendes, R., Alves Silva, H., Anacleto, P., & Cardoso, C. (2011). Effect of CO2 dissolution
Hough, G., Garitta, L., & Gómez, G. (2006). Sensory shelf-life predictions by survival on the shelf life of ready-to-eat Octopus vulgaris. Innovative Food Science and
analysis accelerated storage models. Food Quality and Preference, 17, 468–473. Emerging Technologies, 12, 551–561.
Hough, G., Langohr, K., Gómez, G., & Curia, A. (2003). Survival analysis applied to sen- Mexis, S. F., Chouliara, E., & Kontominas, M. G. (2009). Combined effect of an oxygen
sory shelf life of foods. Journal of Food Science, 68, 359–362. absorber and oregano essential oil on shelf life extension of rainbow trout fillets
Hough, G., Puglieso, M. L., Sánchez, R., & Mendes Da Silva, O. (1999). Sensory and mi- stored at 4 °C. Food Microbiology, 26, 598–605.
crobiological shelf-life of a commercial ricotta cheese. Journal of Dairy Science, 82, Mohapatra, D., Bira, Z. M., Frias, J. M., Kerry, J. P., & Rodrigues, F. A. (2011). Probabilistic
454–459. shelf life assessment of white button mushrooms through sensorial properties
Hough, G., Sánchez, R. H., Garbarini De Pablo, G., Sánchez, R. G., Calderón Villaplana, S., analysis. LWT - Food Science and Technology, 44, 1443–1448.
et al. (2002). Consumer acceptability versus trained sensory panel scores of pow- Montes Villanueva, N. D., & Trindade, M. A. (2010). Estimating sensory shelf life of
dered milk shelf-life defects. Journal of Dairy Science, 85, 1–6. chocolate and carrot cupcakes using acceptance tests. Journal of Sensory Studies,
Hough, G., Wakeling, I., Mucci, A., Chambers, E., IV, Méndez Gallardo, I., & Alves, L. R. 25, 260–279.
(2006). Number of consumers necessary for sensory acceptability tests. Food Qual- Muñoz, A. M., Civille, G. V., & Carr, B. T. (1992). Sensory evaluation in quality control.
ity and Preference, 522–526. New York: Van Nostrand Reinhold.
A. Giménez et al. / Food Research International 49 (2012) 311–325 325

Murray, K. B., & Schlacter, J. L. (1990). The impact of services versus consumers assess- Rozin, P., & Tuorila, H. (1993). Simultaneous and temporal contextual Influences on
ment of perceived risk and variability. Journal of the Academy of Marketing Science, food acceptance. Food Quality and Preference, 4, 11–20.
18, 51–65. Saguy, S., & Peleg, M. (2009). Accelerated and parallel storage in shelf life studies. In H. R.
Nattress, L. A., Ziegler, G. R., Hollender, R., & Peterson, D. G. (2004). Influence of hazel- Moskowitz, S. Saguy, & T. Straus (Eds.), An integrated approach to new food product de-
nut paste on the sensory properties and shelf-life of dark chocolate. Journal of Sen- velopment (pp. 429–455). Boca Raton: CRC Press, Francis and Taylor.
sory Studies, 19, 133–148. Salam, K. I. (2007). Chemical, sensory and shelf life evaluation of sliced salmon treated
Nellman, C., MacDevette, M., Manders, T., Eickhout, B., Svihus, B., Prins, A. G., et al. with salts of organic acids. Food Chemistry, 592–600.
(2009). The environmental food crisis — The environment's role in averting future Salvador, A., Varela, P., & Fiszman, S. M. (2007). Consumer acceptability and shelf life of
food crises. Norway: United Nations Environment Programme (UNEP). “Flor de invierno” pears (Pyrus communis L.) under different storage conditions.
Nielsen, B. R., Stapelfeldt, H., & Skibsted, L. H. (1997). Early prediction of the shelf-life of Journal of Sensory Studies, 22, 243–255.
medium-heat whole milk powders using stepwise multiple regression and princi- Schmidt, K., & Bouma, J. (1992). Estimating shelf-life of cottage cheese using hazard
pal component analysis. International Dairy Journal, 7, 341–348. analysis. Journal of Dairy Science, 75, 2922–2927.
Nunes, M. C. N., Emond, J. P., & Brecht, J. K. (2006). Brief deviations from set point tem- Severson, H. H., Slovic, P., & Hampson, S. (1993). Adolescents' perception of risk: Un-
peratures during normal airport handling operations negatively affect the quality derstanding and preventing high risk behaviour. Advances in Consumer Research,
of papaya (Carica papaya) fruit. Postharvest Biology and Technology, 41, 328–340. 20, 177–182.
Østli, J., Esaiassen, M., Garitta, L., Nøstvold, B., & Hough, G. (in press). How fresh is Silva Sant'Ana, L., Soares, S., & Vaz-Pires, P. (2011). Development of a Quality Index
fresh? Perceptions and experience when buying and consuming fresh cod fillets. Method (QIM) sensory scheme and study of shelf-life of ice-stored blackspot
Food Quality and Preference, http://dx.doi.org/10.1016/j.foodqual.2012.05.008. seabream (Pagellus bogaraveo). LWT — Food Science and Technology, 44, 2253–2259.
OTA (1979). Open shelf-life dating of food. Washington DC: Office of Technology Siripatrawan, U., & Noipha, S. (2012). Active film from chitosan incorporating green tea
Assessment. extract for shelf life extension of pork sausages. Food Hydrocolloids, 27, 102–108.
Patsias, A., Chouliara, I., Badeka, A., Savvaidis, I. N., & Kontominas, M. G. (2006). Stone, H., & Sidel, J. L. (2004). Sensory evaluation practices. London: Elsevier Academic
Shelf-life of a chilled precooked chicken product stored in air and under modified Press.
atmospheres: Microbiological, chemical, sensory attributes. Food Microbiology, 23, Stuart, T. (2009). Waste: Uncovering the global food scandal. London: Penguin Books.
423–429. Symons, H. (2000). Frozen foods. In D. Man, & A. Jones (Eds.), Shelf-life evaluation of
Peleg, M., Normand, M. D., & Corradini, M. G. (2011). Expanded ‘Fermi Solution’ for es- foods (pp. 234). Gaithersburg, Maryland: Aspen Publication.
timating the relationship between product spoilage or deterioration and the num- Varela, P., Salvador, A., & Fiszman, S. (2005). Shelf-life estimation of Fuji apples: Senso-
ber of consumer complaints. Trends in Food Science & Technology, 22, 341–349. ry characteristics and consumer acceptability. Postharvest Biology and Technology,
Peryam, D. R. (1964). Consumer preference evaluation of the storage stability of foods. 38, 18–24.
Food Technology, 18, 214. Vaz-Pires, P., Seixas, P., Mota, M., Lapa-Guimarães, J., Pickova, J., Lindo, A., et al. (2008).
Peryam, D. R., & Pilgrim, F. J. (1957). Hedonic scale method of measuring food prefer- Sensory, microbiological, physical and chemical properties of cuttlefish (Sepia
ences. Food Technology, 11, 9–14. officinalis) and broadtail shortfin squid (Illex coindetii) stored in ice. LWT — Food
Piagentini, A. M., Mendez, J. C., Guemes, D. R., & Pirovani, M. E. (2005). Modelling Science and Technology, 41, 1655–1664.
changes of sensory attributes for individual and mixed fresh-cut leafy vegetables. Ventour, L. (2008). The food we waste: Food waste report v2. UK: WRAP.
Postharvest Biology and Technology, 38, 202–212. Walkling-Ribeiro, M., Noci, F., Cronin, D. A., Lyng, J. G., & Morgan, D. J. (2009). Shelf life
Ramírez, G., Hough, G., & Contarini, A. (2001). Influence of temperature and light expo- and sensory evaluation of orange juice after exposure to thermosonication and
sure on sensory shelf life of a commercial sunflower oil. Journal of Food Quality, 24, pulsed electric fields. Food and Bioproducts Processing, 87, 102–107.
195–204. Wansink, B., & Wright, A. O. (2006). “Best if used by…” How freshness dating influ-
Robertson, G. L. (2006). Food packaging, principles and practices (2nd ed.). Boca Raton, ences food acceptance. Journal of Food Science, 71, 354–357.
FL: CRC/Taylor & Francis. Weber, E. U., & Milliman, R. A. (1997). Perceived risk attitudes: Relating risk perception
Rocha, A. M. C. N., & Morais, A. M. M. B. (2003). Shelf life of minimally processed apple to risky choice. Management Science, 43, 123–144.
(cv. Jonagored) determined by colour changes. Food Control, 14, 13–20. Zhou, T., Harrison, A. D., McKellar, R., Young, J. C., Odumeru, J., Piyasena, P., et al. (2004).
Rodríguez, O., Losada, V., Aubourg, S. P., & Barros-Velázquez, J. (2004). Enhanced Determination of acceptability and shelf life of ready-to-use lettuce by digital
shelf-life of chilled European hake (Merluccius merluccius) stored in slurry ice as image analysis. Food Research International, 37, 875–881.
determined by sensory analysis and assessment of microbiological activity. Food
Research International, 37, 749–757.
Rozin, P. (1990). Social and moral aspects of eating. In I. Rock (Ed.), The legacy of Solo-
mon Asch: Essays in cognition and social psychology (pp. 97–110). Potomac, MD:
Lawrence Erlbaum.

You might also like