COMPOSITIONS OF FUZZY PREFERENCES

 

ANNIBAL PARRACHO SANT’ANNA[1]

               

 

Abstract

 

In this paper a probabilistic methodology to evaluate finite sets of alternatives under multiple criteria is developed. This methodology is based on randomising initial classifications and determining the probability of choice of each alternative as the best according to each criterion. Pessimistic and optimistic points of view are determined. The possibilities of criteria with different importance and statistical correlation between criteria are considered.

 

 Keywords: fuzzy preferences – probability of choice – decision aid

 

 

 

1.    Introduction

Petrovic and Petrovic (2002) developed an approach for multicriteria decision-making that takes into account the optimistic or pessimistic mood that the evaluator desires to assume. Other important feature of their methodology is that not only the evaluation of each alternative according to each criterion but also the weight previously attributed to each criterion may have a linguistic form and a fuzzy representation. The evaluation of the probability of each alternative being chosen according to each criterion plays also an important role in it.

Sant’Anna and Sant’Anna (2001) have developed a different approach that has the same basis on the transformation of initial evaluations into probabilities of choice. Here we extend this approach to provide optimistic and pessimistic extreme evaluations and propose a probabilistic form of combining them. The advantage of this approach is that, these final evaluations being set in strictly probabilistic terms, the composition of the optimist and pessimist value is less susceptible to the influence of scales of measurement. 

The optimistic evaluation is here set in terms of maximizing the probability of choice according to at least one criterion while the pessimistic is set in terms of maximizing the probability of avoiding the worst performance with respect to any criterion. The assumptions underneath this optimistic evaluation are that any strategy to excellence is equally satisfactory and an option is excellent whenever it presents the best results under the criterion that favours it most. On the other hand, the pessimistic idea is that failure to satisfy any criterion will disqualify the effort toward excellence. Besides, while the optimistic evaluation is taken in comparison to the observed options, to the pessimistic evaluation a threshold of worst behaviour with respect to each criterion must also be set. These thresholds may be derived from theoretical limits or from the observed performances.    

In the following section, I describe the procedure adopted to transform the initial evaluations into probabilities of choice. Different approaches to randomisation are there discussed. To provide an example, the modelling approach taken by Sant’Anna and Sant’Anna (2001), based on independent, continuous and uniform membership functions with the same range, derived from the observed ranges, is compared to other approaches to the randomisation of a five options qualitative ranking.

In Section 3, the optimistic and pessimistic composition approaches proposed are applied to Petrovic and Petrovic (2002) example of three alternative options compared under seven criteria. Final evaluations are calculated under the hypothesis of equally important criteria and under a hypothesis of dominance. The results are similar to those obtained by those authors. The possibility of dependence between the criteria is also studied.

Finally, in Section 4, Petrovic and Petrovic (2002) proposal of associating final measures to predetermined intermediate levels of pessimism or optimism is further developed.

 

2.  Fuzzy Preferences and Probabilities of Choice

I believe that the precise form of quantifying preference for a particular option in a set of possible options is through the probability of its choice. But the probability of choice is difficult to access directly. For each preference criterion there is a more natural form of expliciting the relative position of the options: in terms of ranks, in terms of values of numerical variables, such as cost or speed, or in terms of the common language, such as low, moderate or high preference. Preferences presented in linguistic terms suggest representation by fuzzy sets determined through membership functions such as those employed by Petrovic and Petrovic (2002). Randomisation of ranks is also easy to perform (Sant’Anna and Sant’Anna, 2001). Even criteria naturally presented through numerical measures always involve some imprecision. For instance, the preference represented by a precise measurement such as the length of a section road to be built is in fact less deterministic than that measurement. Even if there is no possibility of future changes affecting the length, the inconvenience for the traveller associated to the length, which is what in fact matters and the criterion intends to measure, may be affected by unforeseeable traffic or scenic factors.

We may access the probabilities of choice through a three stages procedure. First obtain a vector of preferences for the available options in the easiest terms. Then transform these punctual measurements into random variables. Finally derive from the distribution of these random variables the probabilities of choice.

            As in the first, different alternatives are also available in the second stage, of modelling the distribution of probabilities of the options evaluations. The initial measurements provide reference points, from which immediately derive the mean of the random distribution. Equalization assumptions may speed the process of modelling the other parameters and simplify the future interpretation of results. Sant’Anna and Sant’Anna (2001) develop assumptions of independence between the distributions representing different options, symmetry around the means and identical dispersion parameters. A uniform continuous distribution is there also proposed as a basic option to be assumed whenever no contrary hints on the distribution form are available. Once a uniform distribution is assumed, determining the range will complete modelling.

Range estimates may also be derived from the initial measurements. The range should be large enough to allow for change of position between the admissible options, so that, if any two production units belong to the set under comparison, there must be a nonnull probability of inversion of their positions. And, since this probability should be small when the considered units are those with the largest and smallest values, we can simplify matters by adding a small parcel to the observed range to generate a common estimate for the range of the distribution of each measure.

 In this article we will adopt the above assumptions and determine the range by adding to the observed range a fraction of 1/(n-1) of it, where n is the number of options. An algorithm to perform the computation of the probabilities of choice under these assumptions is provided in the Appendix.

In Sant’Anna (2002), a comparison between uniform and normal assumptions is presented. The distributional assumptions do not substantially affect the results there obtained. Petrovic and Petrovic (2002) employ discrete, asymmetric and different distributions to represent a classification stated in terms of verbal evaluations. The probabilities of choice may then be derived in the same fashion. We shall consider here the case of 5 options with preferences verbally set as very low, low, moderate, high and very high.

Petrovic and Petrovic (2002) represent these 5 classifications, in a fuzzy set with 5 values, through membership functions of different shapes. In their representation, the graph of the membership function to the random set corresponding to the linguistic classification very high with minimum at 0 and maximum at 1, would be a parable passing through (0,0), (1,1) and (1/2,1/4). Its algebraic expression is given by fvh(x) = x2. To the term very low, symmetrically, the membership function is given by fvl(x) = (1-x)2. To low and high correspond linear membership functions with the same extremes of those corresponding to very low and very high, respectively. Finally the graph of the membership function corresponding to moderate has a triangular shape with vertices at (0,0), (1/2,1) and (1,0). Employing a discrete support set with 5 points, this corresponds to the probability distributions of Table 2.1.

Table 2.1. Five Points Fuzzy Sets for Five Linguistic Ranks

 

lowest

second lowest

median

second highest

highest

very low

8/15

3/10

2/15

1/30

0

low

2/5

3/10

1/5

1/10

0

moderate

0

1/4

1/2

¼

0

high

0

1/10

1/5

3/10

2/5

very high

0

1/30

2/15

3/10

8/15

 

Petrovic and Petrovic (2002) derive probabilities of choice of each of the five options by the minimum of the probabilities of outranking other options. Thus, the probability of the option verbally classified as very high to receive the highest numerical classification is estimated by its probability of obtaining a numerical classification highest than that given to the option with a linguistic classification of high. And the probabilities of choice of the other linguistic classifications are estimated by their probabilities of obtaining a numerical classification highest than that given to the option with the linguistic classification of very high.

            Table 2.2 presents, in the first line, the probabilities of choice derived from Petrovic and Petrovic (2002) approach. The second line presents the probabilities of choice obtained by computing precisely the probabilities of attributing to each fuzzy option the highest numerical values when applying the membership functions of Table 2.1. By looking at the linguistic classification to untie the numerical ties, each of these probabilities may be calculated adding the products of the probabilities of P[r(A) = c], for c varying along the support set, by the product of four summands P[r(B) < c]+P[r(B) = c]1[A > B], obtained by making B assume the four classifications different from A. And in the third line, we present the same probabilities calculated from the randomisation of five equally spaced ranks assuming the uniform distribution as above described and applying the algorithm developed in the Appendix.

Among the three lines of Table 2.2, the last one presents the closest resemblance to the geometric distribution, that has been seen to adequately fit many real practice situations (Lootsma, 1993). This is a possible advantage associated with the uniform model. But the main point in favour of this distribution has to do with its capability of representing ignorance when nothing is known about how possible distortions affect the preferences elicited. A continuous distribution reflects more accurately uncertainty about a given classification than a discrete one and placing a classification as the centre of a uniform distribution is the simplest way to utter a fuzzy preference for this classification.

Table 2.2.  Probabilities of Choice of Five Options

 

 

very low

Low

Moderate

high

very high

probability of passing the highest

4,8%

9,0%

20,8%

57,7%

75,7%

 

probability of being the highest

0,1%

0,4%

1,4%

24,1%

74,0%

 

uniformly being the highest

0,3%

2,5%

10,3%

27,7%

59,1%

 

 

 

3. Probabilistic Composition of Preferences

From the ordering according to different criteria in terms of probabilities of choice we must pass to the global classification of the options. This will be done here by a systematic procedure of composition of the particular probabilities of choice into a global probability. This aggregation procedure must consider at least three different aspects: the relative importance of the criteria, the optimistic or pessimistic global approach to the consequences of the choice and the correlation between the criteria.

 

3.1. The importance of the Criteria

Determining weights to be attributed to the criteria is a much harder task than applying them to classify the concrete options. Simpler than trying to identify, besides values for the options according to the criteria, values for the criteria, as tools to evaluate the same options, is dividing the set of criteria into only two parts: criteria of primary and of secondary importance. I propose here a simple hypercriterion to perform this separation task. The subset of criteria of primary importance will be formed by those criteria that cannot be replaced when computing the probabilities of the options being chosen as the best. The second subset will be formed by those criteria replaceable when considering probabilities of being the best and individually important only when probabilities of avoiding bad performances are computed.

 

3.2. Optimistic and Pessimist points of View

I propose to characterize an optimistic and a pessimistic approach in terms of determining, respectively, optimal or not very bad options. The approach is optimistic when the goal is optimising some performance and pessimistic when what matters is avoiding the proximity of any inferior threshold. The optimistic approach will result in preferring options that present the best indicator according to some criterion. The pessimistic approach will prefer options that do not present any bad evaluation.

Consider, for instance, a criterion given verbally in terms of low, high, very low, very high and moderate satisfaction. If all the alternatives come to be classified in the highly or very highly satisfactory points, from a pessimistic point of view the differences should be given less importance than if the options are classified in the low and very low terms. On the other side, from the optimistic point of view, the difference between high and very high satisfaction should be given much more importance than that given to the difference between low and very low.

The optimistic or pessimistic approach must affect only secondary criteria. Only after the options are classified according to its probability of choice by the primary criteria, we will try to find, in the secondary criteria, some large probability of choice or no large probability of rejection according as an optimistic or pessimistic approach is taken. Supposing, to simplify terms, all criteria presented in an increasing scale from the less desirable to the most desirable situation, the optimistic evaluation will then be put in terms of maximizing the probability of choice according to all the primary criteria and at least one secondary criterion, while the pessimistic evaluation will be put in terms of, besides maximizing all primary criteria, not minimizing any secondary criterion.

If there is no distinction between the criteria, we must attach primary importance to none. Then, in the optimistic approach, we will classify the options by the probability of being chosen according to at least one criterion. Analogously, in the pessimistic approach, we will classify according to the probability of not minimizing any particular probability of choice only.

A pessimist bound will also be added, through a fictitious unit with the lowest admissible position according to each criterion, one step below very low if the criterion is applied in verbal terms.

 

3.3. Correlation

Independence between the criteria should be generally desired, to make easier to evaluate according to them and consequently more reliable the information gathered. But one advantage of the probabilistic approach is that estimates of the correlation can be employed to improve the calculation of the joint probabilities.

A strategy to estimate the probability of joint choices can be based on the knowledge of the correlation coefficient between the choices according to the criteria and the separate probabilities of choice according to each criterion. The correlation coefficient between two Bernoulli random variables is obtained, by definition, dividing the difference between the probability of joint choice and the product of the probabilities of choice according to each criterion by the square root of the product of four terms, these probabilities of choice and their complements. From this we derive an estimate for the joint dependent probability as the sum of the product of the probabilities of choice according to each criterion with this square root multiplied by an estimate of the correlation coefficient.

We combine below probabilistic preferences on roads sections alternatives, assuming independence and, alternatively, assuming a positive correlation between the choices according to length and maintenance cost criteria. Two hypotheses are also considered with respect to the importance attributed to the criteria. First we assume all criteria equally important. Alternatively we apply the methodology to the case of one criterion, pollution, being more important than all the others, as assumed by Petrovic and Petrovic (2002). Computations are also performed under the optimistic and pessimistic points of view above described.

 

3.4. Results

The data set analysed is reproduced in Table 3.1. Length, costs and pollution criteria are to be considered in reverse terms. A last line is added to the original data, presenting the inferior thresholds. Since there are three options, these thresholds were calculated, for the numerical values, by increasing the observed range in one half. For the evaluations in terms from very low to very high, made to correspond to ranks 1 to 5, they were set as zero.

Table 3.2 presents the probabilities of choice according to each criterion and Table 3.3 the probabilities of reaching the lower threshold, under the uniform assumptions of Section 2.

Assuming independence and equal importance to all criteria, option A2 is the best in the pessimistic point of view but option A3 is slightly better if we take the optimistic point of view. If the pollution criterion is considered more important than the others, then option A2 is clearly the best under any approach.

Conceivable correlation levels do not change the classifications. Only if the correlation between length and maintenance cost reaches 0.95, the advantage of option A3 under the optimistic point view with equal importance disappears and option A2 becomes the best under any composition of other assumptions. Table 3.4 presents the final probabilistic classifications under the four different approaches considered, under independence and under correlation of 0.95 between length and maintenance cost.

 

Table 3.1. Evaluation of Three Options according to Seven Criteria

 

length

construction cost

maintenance

speed

security

 impact

pollution

option A1

16,2

156

16

12

high

Moderate

very low

option A2

12,5

151

15

12

high

High

very low

option A3

21,2

128

14

11

very high

High

low

threshold

23,7

170

17

10,5

0

0

0

 

 

Table 3.2. Probabilities of Choice under Uniform Assumptions

 

length

construction cost

maintenance

speed

Security

 impact

pollution

option A1

25%

4%

3%

49%

23%

14%

41%

option A2

72%

10%

22%

49%

23%

43%

41%

option A3

3%

86%

75%

1%

53%

43%

27%

 

 

Table 3.3. Probabilities of Reaching the Thresholds

 

length

construction cost

maintenance

speed

Security

 impact

pollution

option A1

2%

19%

22%

0%

2%

3%

0%

option A2

0%

10%

3%

0%

2%

0%

0%

option A3

30%

0%

0%

22%

0%

0%

2%

 

 

Table 3.4. Evaluations under Different Points of View

 

INDEPENDENT

LENGHT AND MAINTENANCE HIGHLY CORRELATED

 

Equal importance

pollution primary

equal importance

pollution primary

 

OPTIM.

PESSIM.

OPTIM.

PESSIM.

OPTIM.

PESSIM.

OPTIM.

PESSIM.

option A1

86%

59%

32%

24%

88%

54%

33%

22%

option A2

97%

85%

40%

35%

99%

85%

41%

35%

option A3

99%

54%

17%

9%

99%

54%

17%

9%

 

 

 

4.      Weighting Pessimism and Optimism

Petrovic and Petrovic (2002) suggest combining the optimist and pessimist evaluations and derive a pessimistic-optimistic index a from pessimistic-optimistic positions in a scale of five verbal descriptions. The value of a corresponding to each linguistic classification would be the probability of the random number representing this classification appearing as the best in the case of independent assignation of numerical values to the linguistic terms according to predefined membership functions.

The n-th value of a in the scale from the pessimistic to the optimistic extreme would then be given by the probability of the truly n-th best option appearing as the best. The first value, representing the next to maximum optimistic approach, with the identical uniform continuous membership function, would then be 0,74. The value 1-a = 0,26 complementarily corresponds to weight the probability of not being the worst in any criterion by the probability of a classification as very low not generating the lowest numerical rank. A simplest approach would be to consider 5 equally spaced values for a: 0, 0.25, 0.5, 0.75 and 1.

These a values do not take into account the scale in which the final preferences reflecting the two opposite points of view are set. Since our pessimistic approach asks for satisfaction of a series of simultaneous conditions and the optimistic point of view asks for satisfaction of alternative conditions, the first values tend to be smaller than the later. One way to avoid the effect of this difference is to calculate the final probabilities conditionally on the probability that some option satisfy the respective conditions. In the optimistic evaluation, this total probability is always equal to 1, while, in the present case, for the independent equal importance assumption, the pessimistic total is about 97%. Table 4.1, for this hypothesis of independent and equally important criteria, presents the final conditional probabilities under different prior preference levels for the optimistic and pessimistic approaches.

It may also be interesting to determine the value of alpha for which the choice changes, if such a change may occur. In our application, under the hypotheses of independent and equally important criteria, only when the preference for the optimistic point view is given by a value of 95% or more the choice of option A3 would prevail.

Table 4.1. Composition of Conditional Probabilities for Different Alpha Levels

 

0

0,05

0,25

0,5

0,75

0,95

1

option A1

60%

62%

67%

73%

80%

85%

86%

option A2

88%

88%

90%

93%

95%

97%

97%

option A3

55%

58%

66%

77%

88%

97%

99%

 

 

5.      Final Comments

A general approach to take into account uncertainty on evaluations and to compose with different levels of pessimism and optimism was here developed on pure probabilistic terms. Other forms of composition could be explored and other probability distributions might be used to mirror more precise information on criteria importance or on uncertainty. Particularly, the use of discrete random sets associated with the empirical determination of the statistical model for the preferences can be implemented without any changes in the general formulation proposed.

 

References

F. A. Lootsma (1993), Scale Sensitivity in the Multiplicative AHP and SMART, Journal of Multicriteria Decision Analysis, 2, 87-110.

S. Petrovic and R. Petrovic (2002), A New Fuzzy Multi-criteria Methodology for Rankin of Alternatives, International Transactions in Operational Research, 9, 73-84.

A. P. Sant’Anna and L. F. Sant’Anna, Randomisation as a Stage in Criteria Combining. Proceedings of the VII ICIEOM, 248-256, Salvador, BR (2001).

A. P. Sant’Anna (2002), Data Envelopment Analysis of Randomized Ranks, Pesquisa Operacional, to appear.

 

 

APPENDIX

 

Computation of the probabilities of choice assuming independent uniform distributions with the same range.

 

Let x(1),…x(n) denote the observed values in decreasing order and x(n+1) the fictitious smallest value added to the sample. The probabilities will not change if we subtract x(n+1) of every value and divide by the range x(1) – x(n+1). So we can assume without loss of generality x(n+1) = 0, x(1) = 1 and the range equal to 1.

Let M(i) denote the maximum x(i) +1/2 of the random distribution of the (i)-th option evaluated and m(i) its minimum, equal to x(i)-1/2.

By the uniform assumption, the probability that the random value X(i) representing the (i)-th option is the maximum can be computed by integrating from m(i) to M(i) the conditional probability of that happening given X(i). Since the conditional probability of X(i) being the maximum given X(i) = x is the product of the probabilities of X(j) < x, for j ¹1, this integral will be the sum of n integrals, each one corresponding to a subinterval of integration determined by the M(j). In each such subinterval, there is a different number of options with nonnull probability of surpassing X(i).

Consider first i = 1. The first subinterval goes from M(2) to M(1) and along this subinterval the conditional probability is 1. Thus the first summand is M(1)-M(2).

The second summand gives the integral from M(3) to M(2) of the probability of X(2) assuming a value smaller than the integrand x(1). This probability is given by x(1) – m(2) and its integral from M(3) to M(2) is given by M(2)2/2 -M(3)2/2 +(M(2) –M(3))*m(2).

Analogously the third summand is the difference between three sums. The parcels in the minuend of this difference are M(3)3/3, M(3)2/2 multiplied by m(2)+m(3) and M(3) multiplied by m2*m3. The parcels in the subtrahend have the same form with M(3) replaced by M(4).

The same procedure will follow with increasing number of powers, until reaching the lowest subinterval.

For i ¹ 1, the same conditioning approach and the subsequent partitioning works. For i =2, n-1 integrals will have to be computed, since X(2) does not take values in the subinterval (M(2),M(1)). For i = 3, n-2 integrals, and so on.

To speed computation we may also use the fact that the sum of all the products of a number m among the k first m(j) is equal to the sum of all the products of m-1 factors among the first k-1 multiplied by m(k) plus the sum of all products of m factors in the same set of the first k-1.