Sie sind auf Seite 1von 15

Transportation Research Part E 46 (2010) 6175

Contents lists available at ScienceDirect

Transportation Research Part E


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

A consolidated model of trip distribution


Louis de Grange a,*, Enrique Fernndez b,1, Joaqun de Cea b,1
a
b

Department of Industrial Engineering, Diego Portales University, Vergara 432, Santiago, Chile
Department of Transport Engineering and Logistic, Ponticia Universidad Catolica de Chile, Vicua Mackenna 4860, Santiago, Chile

a r t i c l e

i n f o

Article history:
Received 29 December 2008
Received in revised form 8 May 2009
Accepted 2 June 2009

Keywords:
Trip distribution
Multi-objective optimization
Gravity model
Entropy
Maximum likelihood
Spatial aggregation

a b s t r a c t
This work analyzes and compares various trip distribution models with spatial aggregation
within a common theoretical framework for formulating and solving multi-objective optimization problems. A new model is designed that incorporates the main characteristics of
existing ones. These models are then calibrated with a single database at different spatial
aggregation levels using maximum likelihood. The results show that with aggregated data
the various models differ little, but with disaggregated data the differences are considerable. It is also demonstrated that changing the level of data aggregation can signicantly
alter the models parameter values.
2009 Elsevier Ltd. All rights reserved.

1. Introduction
Trip distribution models are intended to produce the best possible predictions of travelers destination choices on the basis of trip generation and attraction information for each travel zone and the level of impedance or generalized cost of traveling between each pair of zones.
The various types of distribution models can be categorized according to whether they are aggregate or disaggregate
(Cascetta et al., 2007). The difference, in the view of some authors, is that the former uses data aggregated over a series
of geographic subdivisions that together represent an activity system served by a transportation network, while the latter
is employed with individual user data and is generally derived from notions of utility theory. Others suggest that utility theory can in fact be applied to aggregated models as well, and point to the disaggregate models segmentation of demand into
behaviorally homogeneous traveler groups and trip purposes as the main distinguishing factor. The present work deals with
models from the aggregate category, many of which can be obtained as the solution to a mathematical programming
problem.
The best-known and most basic aggregate distribution model is the so-called transportation or Hitchcock problem (Hitchcock, 1941) in which goods produced at given origins are supplied to a series of destinations at minimum cost. This model is
formulated as a linear programming problem with constant costs. An improvement on Hitchcock appeared with the development of a doubly constrained adaptation of the classical gravity model (Wilson, 1970). Using the concept of entropy, and
given certain trip generations and attractions, this formulation determines the most likely trip distribution matrix that minimizes total travel costs in a transportation system.
* Corresponding author. Tel.: +56 2 676 8118; fax: +56 2 676 8130.
E-mail addresses: louis.degrange@udp.cl (L. de Grange), jef@puc.cl (E. Fernndez), jdc@puc.cl (J. de Cea).
1
Tel.: +56 2 354 4270 8118; fax: +56 2 553 0281.
1366-5545/$ - see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.tre.2009.06.001

62

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

Many other entropy-based models have been developed for deriving the most likely trip matrix subject to a series of additional exogenous constraints. Among them we may mention Fotheringham (1983, 1986), Fotheringham and OKelly (1989),
Erlander and Stewart (1990), Fang and Tsao (1995), and Thorsen and Gitlesen (1998). These designs are similar to Wilsons
doubly constrained gravity model but with certain sophisticated additions that improve the modeling results. As will be
shown in what follows, they are all obtainable from one specic equivalent optimization problem.
Models based on other techniques such as the intervening opportunities concept of destination choices (Stouffer, 1940;
Schneider, 1959) have also been suggested, but so far they have not proven to be an advance over the entropy-based designs.
We also note that a combined equilibrium model of trip distribution and assignment starting with Wilsons entropy approach and adding the Wardrop equilibrium conditions in the form of a Beckman transformation (Beckman et al., 1956)
has been proposed by Evans (1976). In order to derive combined models comparable to Evans specication, however, we
would require additional assumptions and developments that are outside the scope of the present research.
The remainder of this work is organized as follows. Section 2 analytically deduces a number of entropy-based distribution
models reported in the literature as solutions to the corresponding non-linear optimization problems. Section 3 formulates a
new consolidated model that combines the properties of each of these other models. Section 4 presents a method for estimating the models parameters, describes the main characteristics of the data used for calibration and sets out the main results of the calibration process. A brief statistical analysis is also developed in this section for three levels of data aggregation
corresponding to three different ways of subdividing the Santiago, Chile urban area. Finally, Section 5 sums up the main conclusions and recommendations of our analysis.
2. Formulation of maximum entropy trip distribution models
2.1. Doubly constrained distribution model (GM)
The traditional gravity distribution model (GM) as stated by Wilson (1970) is the following:

T ij Ai Oi Bj Dj ebC ij
1
;
Ai P
bC ij
B
D
j
je
j

1
1
Bj P
bC ij
A
O
i
ie
i

where Cij is the generalized cost of travel between zones i and j (constant in the present case), Tij are the trips between the
two zones, C is the total system cost (constant but unknown), Oi are the trips generated by zone i and Dj the trips attracted by
zone j, Ai and Bj are the typical balancing factors of the doubly constrained gravity model.
Models (1) and (2) are obtained by solving the following optimization problem and applying the rst order conditions:

min
fT ij g

s:t:

Z1
X

T ij ln T ij  1

ij

T ij Oi

li

T ij Dj

cj

X
i

C ij T ij C

ij

The Lagrange multipliers li and cj are the dual variables of the trip generation and attraction constraints. They indicate the
variations in entropy (the objective function) for a unit variation in trip generation or attraction, respectively. The Lagrange
multiplier b expresses the variation in entropy for a unit variation in the total system cost, and can be interpreted as the
travel cost elasticity of demand.
Interestingly, (3) is the substitute form of the following bi-objective problem (Cohon, 1978):

min
fT ij g

min
fT ij g

s:t:

F1

F2

C ij T ij

ij

T ij ln T ij  1

ij

T ij Oi

P
i

T ij Dj

li
cj

min
fT ij g

! s:t:

Z1
P
j

P
ij

C ij T ij 1b

T ij Oi li

P
ij

T ij ln T ij  1
4

T ij Dj cj

Note that if parameter b is large, the relative weight of objective F2 is reduced and the cost-minimization objective F1 is that
much more signicant. If in the limit b tends to innity, (1) will generate the same results as Hitchcocks classical model and
the latter becomes a special case of the doubly constrained gravity model. This relationship was generalized for large populations by Smith (1990) and later validated by Roy and Thill (2004).

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

63

2.2. Distribution model based on competing destinations (CDM)


The second model we consider is the competing destinations model (CDM). Developed by Fotheringham (1983, 1986), it
takes the following functional form:

T ij Ai Oi Bj Dj Sij q ebC ij
1
Ai P
q bC ij ;
B
D
S
j j j ij e

5
1
Bj P
q bC ij
A
O
S
i i i ij e

where Sij is the attractiveness of traveling between zone i and zone j considering the accessibility of destination j. Thus, Sij is
the accessibility of destination j relative to all other destinations as perceived from the standpoint of i. Fotheringham denes
Sij as

Sij

w
X

Dk erCjk

or Sij

k1
ki;kj

w
X

Dk C jk r

k1
ki;kj

where parameter r is a weighting factor representing the impedance of traveling between i and j and variable w is the number of potential destinations. Note that denition (7) transforms the problem into a non-linear model in the parameters that
may be very complex to estimate and does not guarantee a unique solution. For this reason, Thorsen and Gitlesen (1998),
Guldmann (1999) and others dene r = 1.
The sign of parameter q in (5) is determined empirically by the particular model implemented. It is the result of two
forces dened by Fotheringham that impact on the choice of destination:
(i) The agglomeration force, related to the impedance or cost of traveling between two zones. As this cost rises, the number of trips between the zones tends to fall. The agglomeration force is normally incorporated into all distribution
models, particularly those based on entropy analysis.
(ii) The competition force, related to the fact that the greater is the distance a traveler is willing to travel, the more destination alternatives there are and the higher, therefore, is the probability of trip satisfaction in terms of increased
access to employment, shopping or education options.
If empirical analysis reveals that q > 0, the agglomeration force is dominant, while if q < 0, it is the competition force that
dominates.
Model (5) could be obtained directly by solving the following multi-objective optimization problem:

min F 1
fT ij g

T ij C ij

T ij ln T ij  1

ij

min F 2
fT ij g

X
ij

max F 3
fT ij g

s:t:

T ij lnSij

10

ij

T ij Oi

8i li

11

T ij Dj

8j cj

12

X
i

Thus, we observe that Fotheringhams model is a classical Wilson-type gravity model but with the addition of an objective
(10) that maximizes the natural logarithm of the attractiveness of travel between zones i and j for all such pairs in the system. A substitute version of the problem is

min Z 2
fT ij g

s:t:

T ij C ij

ij

1X
qX
T ij ln T ij  1 
T ij ln Sij
b ij
b ij

T ij Oi

li

T ij Dj

kj

13

X
i

where parameters b and q are the relative weights of objectives (8) and (10), respectively. By solving problem (13) we obtain
the optimality conditions (5) and (6) of the Fotheringham model. Note that objective (10) could be dened as a restriction to
the problem, and q/b as its respective shadow price. In such case, this restriction represents the minimum level of attraction
in the system that has to be satised.

64

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

An interesting extension of Fotheringham, given by Thorsen and Gitlesen (1998), assumes the following functional form:

T ij Ai Oi Bj Dj Oi a Dj g Sij q ebC ij
Ai

Oi a

1
;
Bj Dj Dj g Sij q ebCij

14
Bj

Dj g

1
Ai Oi Oi a Sij q ebC ij

15

Eq. (14) exhibits a higher degree of exibility than (5) in that the ratio of trips between OD pairs to both trip generations
and trip attractions is not strictly linear (except when a = 0 and g = 0). The inclusion of the term (Oi)a(Dj)g allows this model
to incorporate a greater degree of spatial heterogeneity, but it must be validated empirically. From the standpoint of parameter estimation, model (14) has fewer degrees of freedom than model (5) and should also afford a better statistical t. However, in our calibration process (Section 4 below) parameters a and g were not included in the model as their statistical
signicance (critical values) were close to zero.
2.3. Self-deterrent distribution model with quadratic costs (SDM)
The third model in our analysis is an entropy trip distribution model with quadratic costs (Fang and Tsao, 1995).

T ij Ai Oi Bj Dj ebC ij kT ij C ij
Ai P

1
;
Bj Dj ebC ij kT ij Cij

16
Bj P

1
Ai Oi ebC ij kT ij Cij

17

Known as the self-deterrent gravity model (SDM) by its developers (16), is in fact similar to the classic gravity model (1), the
difference being the inclusion of variable Tij in the exponent. The congestion term kCijTij in (16) must be validated during the
parameter estimation process. More specically, given that Tij > 0 and Cij > 0 by construction, the presence of this term requires that b and k carry the same sign. And since they are both preceded by negative signs in the exponent, their estimated
values must be positive.
Fang and Tsao (1995) did not estimate the two parameters, conning themselves to simulations with values whose signs
were the same (congestion case, e.g. private transport). If the sign of k were in fact the opposite of bs, there would be economies rather than diseconomies of scale (e.g. public transport). Such a situation could occur for certain trips whose attractiveness increased when more travelers chose it. Since as noted, Fang and Tsao only provide simulations for certain values,
actual estimations of these parameters were conducted for this study and are will be set out here in Section 4.
Models (16) and (17) are derived by solving the following optimization problem:

min Z 3
fT ij g

s:t:

T ij C ij

ij

1X
k X 2
T ij ln T ij  1
T C ij
b ij
2b ij ij

T ij Oi li

18

T ij Dj cj

P
By including the quadratic term ij T 2ij C ij the authors introduce non-linearities in the cost structure of the model to accommodate the concept of congestion. As the number of trips between pair (i, j) increases the cost of the trip will also rise, the
formulation thus exhibiting diseconomies of scale.
The problem (18) from which SDM is derived is a substitute for the following multi-objective optimization problem:

min F 1
fT ij g

T ij C ij

19

T ij ln T ij  1

20

C ij T 2ij

21

ij

min F 2
fT ij g

X
ij

min F 4
fT ij g

s:t:

X
ij

T ij Oi

li

22

T ij Dj

cj

23

X
i

Observe that like Fotheringham, this model is also a classical Wilson-type gravity model with an added objective (21), which
in this case minimizes the sum of the squares of the trips weighted by their respective costs. Indeed, the additional objective
is the equivalent of objective function in the model proposed by Morrison and Thumann (1980) with weight parameters pij
set to 1/Cij (see details in Appendix A). This approach confers more degrees of freedom on the model and better captures the

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

65

non-linearity of the impact of costs on trip demand. Note that since (16) is a xed-point equation, complex numerical methods have to be used to obtain a solution (Fang and Tsao, 1995). On the other hand, the calibration process employed here
(Section 4) does not present any additional complexity.
2.4. Distribution models incorporating an a priori matrix
If observed trip information is available from an a priori matrix obtained, for example, from a ridership survey, such data
can be used in the modeling process. Including observed trips between the (i, j) pair, which we will call Nij, will modify the
entropy equation present in almost all of the models surveyed here so far.
To demonstrate how this matrix data is introduced, consider the Wilson (1970) model given above in (3). Upon inclusion
of this information the modied version of the equivalent optimization problem takes the following form (Willumsen, 1978;
Van Zuylen and Willumsen, 1980):

0
1T ij
T! Y B Nij C
max F Q
@P A
fT ij g
T ij ! ij
Nij
X

s:t:

ij

ij

T ij Oi

li

T ij Dj

cj

24

X
i

C ij T ij C

ij

P
The term N ij = ij N ij represents the proportion of trips in the survey sample between the (i, j) pair. Therefore, the expression

T ij
PNij
represents the joint probability of making Tij trips between i and j. Using Stirlings approximation to handle the
ij

Nij

factorial expression in (24), and assuming T is constant, we obtain the following equivalent optimization problem:

min F

XX

fT ij g

s:t:

T ij ln T ij  1 

XX

T ij Oi

li

T ij Dj

cj

T ij ln N ij  ln

!
Nij

ij

25

X
i
X

C ij T ij C

ij

Upon applying the optimality conditions in (25) we have

T ij Ai Oi Bj Dj ebC ij pij
Ai P

1
;
Bj Dj ebC ij pij

26
1
Bj P
Ai Oi ebC ij pij

27

P
where pij N ij = ij N ij . Thus, if there are no trips in the matrix between a given pair (k, l), the number of trips predicted by the
model (26) will also be zero for that pair; that is (Tkl = 0) given that Nkl and therefore pkl = 0.
With the incorporation of an a priori matrix as in (26) we can make estimates only for the cells that have trip information.
This considerably limits the use of the distribution model for generating trip predictions given that in real-life situations,
such as a major city with large numbers of origins and destinations, more than 95% of the cells in the trip matrix may be
empty.
However, if calibration is done using maximum likelihood, one of the most widely used techniques for estimating distribution model parameters (see Flowerdew and Aitken, 1982; Sen, 1986; Sen and Matuszewski, 1991; Congdon, 1992; WinP
kelmann and Zimmermann, 1995; Hu and Pooler, 2002), the presence of the pij N ij = ij N ij term will have no impact since
only those (i, j) pairs for which there are trips in the sample will be utilized in the likelihood function.
Returning to our objective function (25), this formulation can be rewritten as follows (Willumsen, 1978, 1981,1984; Van
Zuylen and Willumsen, 1980; Ortzar and Willumsen, 2001):

XX
i

1


X
T ij
B
C XX

T ij @ln T ij  ln N ij 1  ln
N
T
ln

const
A
ij
ij ij
Nij
|{z}
i
j
const

28

66

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

Considering that

lnx




2

3
x1
1 x1
1 x1



x
2
x
3
x

T ij
N ij

and dening

29

x, (28) can be approximated as

XX 1
T ij  Nij 2
Nij
i
j

30

This expression is just the objective function of the distribution problem based on squared differences (Morrison and Thumann, 1980), already presented here as Appendix B.
3. Formulation of a maximum entropy consolidated model (CM)
We now present a new trip distribution formulation we shall call the consolidated model (CM) that brings together the
most important characteristics of the three entropy models described in the previous section. After stating the multi-objective optimization problem, we give the corresponding substitute problem and then apply the optimality conditions to obtain
the model itself.
The multi-objective problem incorporating the key properties of the three models has the following structure:

min F 1
fT ij g

X
ij

min F 2

T ij C ij

fT ij g

31

T ij ln T ij  1

32

T ij lnSij

33

ij

max F 3
fT ij g

min F 4
fT ij g

s:t:

ij

C ij T 2ij

34

ij

T ij Oi

li

35

T ij Dj

cj

36

X
i

Objective (31) is taken from the Hitchcock model, objective (32) from the Wilson model, objective (33) from the Fotheringham model and objective (34) from Fang and Tsao. A substitute for this problem is:

min Z 4
fT ij g

s:t:

T ij C ij

ij

1X
qX
k X
T ij ln T ij  1 
T ij ln Sij
C ij T 2ij
b ij
2b ij
b ij

T ij Oi li

37

T ij Dj cj

Applying the optimality conditions to (37) we obtain the following:

T ij Ai Oi Bj Dj Sij q ebCij kCij T ij


1
Ai P
Bj Dj Sij q ebC ij kCij T ij

38
39

Bj P

1
Ai Oi Sij q ebCij kCij T ij

40

To check the uniqueness of the solution (a global minimum) we examine the second-order conditions of the objective function in (37):

@2 Z4
@T 2ij



1
kC ij
1 1

kC ij > 0
>0!
bT ij
b T ij
b

By construction, Tij > 0 and Cij > 0. Thus, to satisfy (41) it must be the case that b > 0 and

41


1
T ij


kC ij > 0. Observe that in (38),

the parameter b carries a negative sign. This implies that when the cost of a trip between pair (i, j) increases, the demand Tij


for such trips declines, which is conceptually correct. In addition, if it is true that T1ij kC ij > 0, then T1ij > kC ij . Since Ti > 0

67

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

and Cij > 0, a positive value of k is sufcient to ensure a unique solution. This is consistent with the effect of congestion on trip
cost between pair (i, j) in that the kTijCij term in (38) is preceded by a negative sign. Note that these uniqueness conditions are
also valid for SDM. As for GM and CDM, uniqueness requires that only Tij be greater than zero, which obviously is always true
(the number of trips cannot be negative).
We now turn to the calibration of the parameters, which is the same process for all of the models considered.
4. Parameter estimation of the trip distribution model
4.1. Parameter estimation
Estimating the gravity model parameters is a subject of frequent debate among researchers and practitioners, and a wide
variety of calibration techniques have been suggested. The two main approaches are based either on statistical and econometric methods (see, for example, Abrahamsson and Lundqvist, 1999; Ortzar and Willumsen, 2001; Ham et al., 2005), or on
formulating and solving non-linear mathematical programming problems (Wilson, 1970; Morrison and Thumann, 1980;
Fang and Tsao, 1995; Gonalves and Souza, 2001).
Since the number of persons traveling between a given OD pair (i, j) is a non-negative integer variable, the Poisson distribution would seem to be a natural choice for the estimation process (see Flowerdew and Aitken, 1982; Sen, 1986; Sen and
Matuszewski, 1991; Congdon, 1992; Winkelmann and Zimmermann, 1995; Hu and Pooler, 2002). If we assume, then, that
the matrix data Nij are independent, Poisson-distributed observations, their probability function is

PNij =T ij

eT ij  T ij Nij
Nij !

42

In the case of our CM model in (38), probability function (42) becomes

PNij =T ij

eAi Oi Bj Dj Sij

q ebC ij kC ij T ij

 Ai Oi Bj Dj Sij q ebC ij kC ij T ij Nij


Nij !

43

Since the likelihood function is built on a sample (the a priori matrix), (43) reduces to

PNij =T ij

eAi Oi Bj Dj Sij

q ebC ij kC ij Nij

 Ai Oi Bj Dj Sij q ebC ij kC ij Nij Nij


Nij !

44

We can then formulate the following log-likelihood function:

max ln L

X X

fb;q;A;Bg

Finally, given that

max ln L

Ai Oi Bj Dj Sij q ebCij kCij Nij Nij ln Ai ln Bi q ln Sij  bC ij  kC ij Nij

45

ij T ij

X X

fb;q;k;A;Bg

ij N ij

T is constant, solving (45) is equivalent to


Nij ln Ai ln Bi q ln Sij  bC ij  kC ij Nij

46

The Poisson model allows us to explicitly include zero values in the modeling without producing biased estimators for long
trips, as would happen with other specications (Sen, 1986). According to Winkelmann and Zimmermann (1995), the model
has the specic merits of capturing the discrete and non-negative nature of the data and permitting inferences to be drawn
about the probability of event occurrence. It also allows for a straightforward treatment of zeros in a multiplicative model.
Note also that on the basis of (38), we can dene the following logit probability function:

T ij
Ai Oi Bj Dj Sij q ebC ij kC ij T ij
eli kj bC ij q ln Sij kC ij T ij
Pij P P
q bC ij kC ij T ij P li kj bC ij q ln Sij kC ij T ij
T ij
Ai Oi Bj Dj Sij e
e
ij

ij

ij

This function represents the probability of traveling between pair (i, j). Since
lowing log-likelihood function (Abrahamsson and Lundqvist, 1999):

max ln L

fb;q;k;A;Bg

47

XX
i

Nij ln Ai ln Bi q ln Sij  bC ij  kC ij N ij 

ij T ij

is constant, we can then dene the fol-

48

Observe that (48) is identical to (46). We may therefore conclude that in the context of the gravity model, a logit approach to
estimating these models parameters is equivalent to using a Poisson distribution.
4.2. The data
The trip information database for estimating the parameters of the various models consisted of bus trip matrices containing data originally compiled from a massive survey of Santiago bus users conducted in 2001 (SECTRA, 2001). The survey was

68

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

conducted on the buses themselves, with three riders chosen at random on every second bus. The set of data we used was
drawn from a matrix for the Santiago areas morning rush hour that was estimated on the basis of a sample of 87,446 riders
constituting 16.67% of the transit system universe. The travel cost (Cij) for each OD pair is the cost for the fastest bus route
(walking time + waiting time + travel time) between the centroid of origin to the centroid of destination. In zones with denser bus networks, the walking and waiting times are lower.
A map of the Santiago area covered by the data is displayed in Fig. 1. As can be seen, three different levels of area subdivision were employed: 577 zones, 36 districts and 7 sectors. Each of these levels corresponds to a different data aggregation level. Various trip matrix characteristics for the three levels are given in Table 1.
For the aggregated cases (i.e., districts and sectors), the generalized cost levels must be calculated using average or aggregated values of Cij for all users of each (i, j) pair; that is, we must estimate a representative Ci0 j0 for each aggregated pair (i0 , j0 ).
Note that (i, j) 2 (i0 , j0 ), where (i, j) is the disaggregated pair and (i0 , j0 ) the aggregated one. In the process of aggregation, Cij for
the 36 districts and 7 sectors (see Section 4.3) was taken as the average cost dened in the initial aggregation (577 zones)
according to the following formula:

C i0 j 0

N ij C ij =Ni0 j0

49

i;j2i0 ;j0

where Nij are the observed trips between pair (i, j) at the disaggregated level, Ni0 j0 are the trips between pair (i0 , j0 ) and
P
0 0
0 0
i;j2i0;j0 N ij N i0j0 . Thus, from the values of Cij and Nij at the disaggregated level we obtain the Ci j and Ni j values for calibrating the aggregated models.
4.3. Analysis of results
4.3.1. Results obtained for the zone aggregation level (577  577)
The results obtained with our four distribution models using the calibration procedure described above (Section 4.1) at
the zone aggregation level (577 zones) are presented in Table 2. The statistics utilized were r2 (correlation between observed
and modeled trips) and Log-L (log-likelihood).
Another indicator of the predictive capacity of distribution models is the standardized root mean square error (SRMSE).
According to Knudsen and Fotheringham (1986), SRMSE is the most accurate measure for analyzing the performance of two
or more models in replicating the same data set, or for comparing a models performance in different spatial systems. In
terms of our models, SRMSE is given by the following expression:

Area Subdivision Levels


577 ZONES

36 DISTRICTS

7 SECTORS

Fig. 1. Area subdivision levels.

Table 1
Bus trip matrix for morning rush hour by level of aggregation, Santiago, 2001.
Characteristic

Zone aggregation

District aggregation

Sector aggregation

No. of area subdivisions


No. of OD pairs
Total trips
No. of cells with trips
No. of cells without trips
Percentage of zeros

577
332,929
524,674
50,151
282,778
84.9%

36
1296
524,674
1201
95
7.3%

7
49
524,674
49
0
0%

69

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175


Table 2
Parameters estimated by maximum likelihood, zone aggregation (577  577).
Model

r2

Log-L

SRMSE

GM
CDM
SDM
CM

0.521
0.539
0.586
0.587

14,796.1
14,733.9
14,726.9
14716.9

3.031
2.988
2.727
2.727

0.279
0.303
0.287
0.331

(227.6)
(129.4)
(237.5)
(142.2)

0.306 (16.9)

0.309 (15.6)

0.00196 (75.8)
0.00172 (68.1)

v
uP
u T ij  Nij 2 , P T ij
u
ij
t ij
IJ
IJ
As can be seen in Table 2, the signs of the parameters were consistent with our theoretical exposition. In addition, for all of
the models the parameters were signicantly different from zero and the r2 values were also acceptable, the CM performing
best on the latter criterion. CM also displayed the lowest value for log-likelihood.
Since GM is just CDM with q = 0, or SDM if k = 0, this implies in econometric terms that the restricted model (GM) can be
compared pairwise with CDM and the SDM. On similar grounds, CDM and SDM may be compared with CM. We can thus
statistically determine whether each of these pairs of models are different or not. To make these pairwise comparisons
we tested the null hypothesis that two models are equivalent by performing the following likelihood-ratio test (Green,
h  v2p , where ^
hR is the estimator of the restricted model parameters, ^
h is the estimator of the non-re2003): 2ln ^
hR  ln ^
stricted model parameters and p is the number of constraints imposed on the restricted model. In the present case, p = 1,
yielding the following results at a 99% condence level:





h 124:41 > 6:63


GM versus CDM: 2ln ^
hR  ln ^
h 138:43 > 6:63
GM versus SDM: 2ln ^
hR  ln ^
h 33:93 > 6:63
CDM versus CM: 2ln ^
hR  ln ^
h 19:91 > 6:63
SDM versus CM: 2ln ^
hR  ln ^

We therefore conclude that at a highly disaggregated level, the various models are not equivalent and that CM is the best
model while GM is the worst.
Fig. 2 displays a comparison of the modeled and observed trips. As can be seen, GM tended to signicantly overestimate
short trips whereas both CDM and CM estimated them more accurately. In the case of distances of less than 1 km, for which
the observed number of trips is 30,000, CM generated a correct prediction while GM produced a gure of almost 68,000. The
CDM estimate of less than 39,000 trips was plainly closer to the actual number than GM, and SDM was considerably further
off at nearly 50,000.

Histogram of Modeled and Observed Trips, Zone Aggregation (577x577)


T-OBS

T-GM

T-CDM

T-SDM

T-CM

70.000

50.000
40.000
30.000
20.000
10.000
0
0-1
1-2
2-3
3-4
4-5
5-6
6-7
7-8
8-9
9-10
10-11
11-12
12-13
13-14
14-15
15-16
16-17
17-18
18-19
19-20
20-21
21-22
22-23
23-24
24-25
25-26
26-27
27-28
28-29
29-30
30-31
31-32
>32

Number of Trips

60.000

Distance (km)
Fig. 2. Histogram of modeled and observed trips, zone aggregation (577  577).

70

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

4.3.2. Results obtained for the district aggregation level (36  36)
Table 3 reports the results of the four distribution models using the same calibration process but with the data aggregated
at the district level.
As was true for the zone aggregation case, the signs of the parameters with district aggregation were consistent with our
theoretical exposition. For all of the models the parameters were signicantly different from zero and the r2 values were also
acceptable, the CM again performing best on the latter criterion. The likelihood-ratio test rejected the null hypothesis that the
models were equivalent:





h 21:63 > 6:63


GM versus CDM: 2ln ^
hR  ln ^
h 22:49 > 6:63
hR  ln ^
GM versus SDM: 2ln ^
h 9:86 > 6:63
CDM versus CM: 2ln ^
hR  ln ^
h 9:00 > 6:63
SDM versus CM: 2ln ^
hR  ln ^

In Fig. 3 we again observe that GM and the SDM overestimated the number of short trips while the CDM and CM predictions
corrected this divergence.
4.3.3. Results obtained for the sector aggregation level (7  7)
Table 4 reports the results for the four distribution models using the same calibration process with sector aggregation, the
last of the three aggregation levels.
Table 3
Parameters estimated by maximum likelihood, district aggregation (36  36).
MODEL

r2

Log-L

SRMSE

GM
CDM
SDM
CM

0.766
0.772
0.775
0.782

2854.6
2843.8
2843.3
2838.8

1.413
1.392
1.399
1.382

0.319
0.351
0.306
0.340

(62.3)
(28.4)
(60.7)
(26.9)

0.560 (2.7)

0.543 (3.1)

0.00003 (13.7)
0.00002 (14.0)

Histogram of Modeled and Observed Trips, District Aggregation (36x36)


T-OBS

T-GM

T-CDM

T-SDM

T-CM

160.000
140.000

Number of Trips

120.000
100.000
80.000
60.000
40.000

>30

27-30

24-27

21-24

18-21

15-18

12-15

9-12

6-9

3-6

0-3

20.000

Distance (km)
Fig. 3. Histogram of modeled and observed trips, district aggregation (36  36).

Table 4
Parameters estimated by maximum likelihood, sector aggregation (7  7).
Model

r2

Log-L

SRMSE

GM
CDM
SDM
CM

0.944
0.946
0.947
0.957

74.7
73.1
73.0
72.5

0.387
0.379
0.379
0.376

0.269
0.214
0.270
0.223

(19.1)
(4.4)
(20.1)
(4.6)

0.771 (1.6)

0.581 (1.1)

0.00000177 (2.5)
0.00000157 (2.2)

71

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

Histogram of Modeled and Observed Trips, Sector Aggregation (7x7)


T-OBS

T-GM

T-CDM

T-SDM

T-CM

250.000

Number of Trips

200.000

150.000

100.000

50.000

>24

21-24

18-21

15-18

12-15

9-12

6-9

3-6

0-3

Distance (km)
Fig. 4. Histogram of modeled and observed trips, sector aggregation (7  7).

Interestingly, at this level the estimator for the parameter q was positive, unlike the other two cases. This implies that,
following the denitions in Fotheringham (1983), the agglomeration forces are dominant when aggregation is high while the
competition forces dominate when it is low.
This difference in sign can be explained via a simple example. When the area subdivisions are relatively large, they embrace a high number of activities and potential destinations related to education, commercial, medical and other services,
and the likelihood travelers need to choose a different or more distant subdivision is relatively low. In other words, the number of destinations that compete to satisfy the users motive for traveling (work, shopping, health care, etc.) is relatively
small. By contrast, when the subdivisions are relatively small, the range of activities is also small and the likelihood travelers
will choose a different subdivision is relatively high. In this case, the number of destinations competing to satisfy the users
motive for traveling is relatively large.
We also observe regarding q that for CDM it was signicant at the 89% condence level (t-test equals 1.1), implying that
the elimination of this parameter is debatable.
On the other hand, as is shown below the likelihood-ratio test at this aggregation level did not reject the null hypothesis
that the models are equivalent, and consequently they all reduce to the gravity model. This being the case, the fact that they
all satisfactorily predicted the observed trips, as can be seen in Fig. 4, conrms what we would intuitively expect. Also, at this
highly aggregated level GM produces a reasonable t, in contrast to its performance at the relatively disaggregated levels.





h 3:19 < 6:63


GM versus CDM: 2ln ^
hR  ln ^
h 3:27 < 6:63
GM versus SDM: 2ln ^
hR  ln ^
h 1:14 < 6:63
CDM versus CM: 2ln ^
hR  ln ^
h 1:07 < 6:63
SDM versus CM: 2ln ^
hR  ln ^

4.3.4. Summary of results


Table 5 summarizes the outcomes of the maximum likelihood estimates for the three levels of aggregation just analyzed.
These results may be summed up as follows:
(i) Parameters change and their statistical signicance declines as the subdivisions of the area under analysis are increasingly aggregated. Thus, the statistical inferences depend on the aggregation level.
(ii) At all aggregation levels the statistical t of the consolidated model is better than the others, while that of the gravity
model is always inferior. This implies that CMs ability to reproduce eld data is also greater. Its superiority is clearly
attributable to the greater number of explanatory variables given that in each maximum likelihood estimation, adding
new explanatory variables increases (or at least maintains) the value of Log-L. As regards the competing destinations
and self-deterrent gravity models, the results are not conclusive at any aggregation level; nevertheless, they are
always superior to GM and inferior to CM.

72

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

Table 5
Parameters estimated by maximum likelihood, all aggregation levels.
Subdivisions

Model

r2

SRMSE

577

Gravity
Competing destinations
Self-deterrent
Consolidated
Gravity
Competing destinations
Self-deterrent
Consolidated
Gravity
Competing destinations
Self-deterrent
Consolidated

0.521
0.539
0.586
0.587
0.766
0.772
0.775
0.782
0.956
0.956
0.971
0.973

3.031
2.988
2.727
2.727
1.413
1.392
1.399
1.382
0.387
0.379
0.379
0.376

0.279
0.303
0.287
0.331
0.319
0.351
0.306
0.340
0.269
0.214
0.270
0.223

36

q
(227.6)
(129.4)
(237.5)
(142.2)
(62.3)
(28.4)
(60.7)
(26.9)
(19.1)
(4.4)
(20.1)
(4.6)

0.306

0.309

0.560

0.543

0.771

0.581

k
(16.9)
(15.6)
(2.7)
(3.1)
(1.6)
(1.1)

0.00196 (75.8)
0.00172 (68.1)

0.00003 (13.7)
0.00002 (14.0)

0.00000177 (2.5)
0.00000157 (2.2)

(iii) At more highly aggregated levels (fewer area subdivisions), all models provide a better statistical t (r2 and SMRSE)
and reproduce the observed trips more accurately than at more disaggregated levels. This is because with greater
aggregation, the number of zone pair trips is lower than at the less aggregated levels and fewer variables thus need
to be estimated.
(iv) As the aggregation level increases, the difference between the four models declines, and vice versa. This is because
with high aggregation (few area subdivisions), trips within the short distance ranges where the models diverge considerably cannot be distinguished. For example, a 06 km range would not differentiate between trips of 1 km and
those of, say, 5 km. Furthermore, with high aggregation the trip distribution tends to be strongly dominated by the
origin and destination constraints. If, on the other hand, aggregation is low (many area subdivisions), the origin
and destination constraints are less important. Note also regarding this last point that as the degree of aggregation
of the total area under analysis declines, the number of OD pairs for which trips must be estimated increases quadratically while the origin and destination constraints do so only linearly.
(v) The higher is the level of disaggregation, the more GM generates distortions with respect to the observed values, primarily in the short trips. The models that include a competing destinations term (CDM and CM) tend to correct these
short-distance distortions. SDM, on the other hand, does not correct the problem of excessive short trips.
(vi) As the aggregation level increases, the competing destination force tends to disappear. This is corroborated by the qparameter values.
(vii) The sign of the k parameter estimated by the SDM and CM models would be consistent with the formulation of the
model, which penalizes (due to congestion) the generalized cost of traveling between pairs (i, j) that register a higher
number of trips.
5. Conclusions
A consolidated approach to the formulation and calibration of entropy trip distribution models was presented that is
based on the specication and solution of multi-objective mathematical programming problems as well as traditional econometric techniques. The main contribution of this work to the literature is the denition of a common theoretical framework
for formulating trip distribution models with spatial aggregation, which was used to specify a new, consolidated model that
incorporated the main characteristics of models previously reported in the literature. Three such existing models as well as
the proposed consolidated one were calibrated using real-world data and comparisons of the four designs were then conducted. The results demonstrated that the traditional gravity model provides the worst statistical t while our proposed version, which combines the features of the other three, exhibited the best t.
A further contribution of our research consists in the nding that the data aggregation level used in calibrating these
models inuenced the parameter estimation results and is thus an important factor in choosing the most appropriate type
of model. More specically, with disaggregated data the most complex formulation (the consolidated model) provided better
goodness-of-t indicators than the simpler (gravity) model, which generated relatively inferior ones. With aggregated data,
however, the various models provided almost the same goodness-of-t outcomes. In situations where disaggregated information is available, therefore, using more complex models is advisable, whereas if the available information is aggregated,
simpler models would be called for as the estimates will be equivalent to those of the complex models.
Our results also demonstrated that the level of data aggregation can impact the trip distribution parameters. Increasing
spatial aggregation (fewer and larger subdivisions) can produce signicant changes in the parameter values and even change
their sign, so care must be taken when giving them an economic interpretation. On the other hand, when the level of disaggregation is high (more and smaller subdivisions), the parameter estimates vary within stable value ranges. In short, different levels of data aggregation can result in distribution models with very different parameters, particularly when
aggregation is high. It is important to recognize that the levels of aggregation normally used in gravity models, for the estimation of trips, corresponds to census zones (577 in our case).

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

73

An obvious avenue for future research would be to further validate the models with data for other cities so that the results
obtained here could be generalized. Another extension of the present study on which the authors are currently working is
the incorporation of the consolidated model into the context of combined network equilibrium models that include the
mode choice and trip assignment steps and accommodate spatial correlation in the calibration process (De Cea et al., 2008).
Appendix A. Derivation of doubly constrained gravity distribution model
To understand the concept of entropy applied to spatial urban trip modeling and the economic interpretation of mathematical programming problems that give rise to spatial distribution trip models, consider the following example.
A transportation system consists of 1 origin generating four trips, and three destinations j. The four trips are distributed
among the destinations as follows: one trip to Destination 1, one trip to Destination 2 and two trips to Destination 3. Since
there are four trips, there are also four travelers (A, B, C and D). The possible ways in which the travelers can make these trips
are listed in Table A.1.
As can be observed, in this example there are 12 possible ways N of distributing the four travelers among four destinations j. The same result can be derived mathematically by expressing the example as the following combinatorial problem:

     
4
3
2
4!
3!
2!





12
1!  3! 1!  2! 2!  0!
1
1
2

A:1

In general, for n destinations the number of ways N that trips Tj may be distributed among the destinations j can be written
as follows:

T!
T!

T 1 !T 2 !T 3 !    Pj T j !

A:2

So far, however, we have only considered the special case where there is a single origin. For the general case in which there
are multiple origins and destinations, the formula is

T!

Pi Pj T ij !

T!

A:3

Pij T ij !

If the travelers are assumed to be homogeneous (as is typically done in aggregate spatial distribution models), the most
probable trip distribution is the one with the largest value of N. The value of N is at its maximum when the distribution
is uniform and all cells in the trip matrix have the same value (in the absence of trip generation and attraction constraints),
and at its minimum when the value for all cells but one is 0, the single exception having a value of T.
In our example, each specic trip distribution for travelers A, B, C and D is known as a microstate, while the distribution as
a whole is the macrostate (a set of microstates).
If we assume that all microstates are equally probable, the spatial trip distribution matrix will be given by the macrostate
that contains the greatest number of associated microstates. The estimated distribution (that is, the one with the highest
probability) will thus be the one that maximizes the value of N (the number of microstates). Therefore, to determine the
most likely matrix we must solve the following optimization problem:

T!
max F
fT ij g
Pij T ij !
X
T ij Oi
s:t:

A:4

T ij Dj

Table A.1
Possible ways of distributing four travelers among four destinations.

1
2
3
4
5
6
7
8
9
10
11
12

j=1

j=2

j=3

A
A
A
B
B
B
C
C
C
D
D
D

B
C
D
A
C
D
A
B
D
A
B
C

CD
BD
BC
CD
AD
AC
BD
AD
AB
BC
AC
AB

74

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

where Tij are the trips between zone i and zone j, Oi are the trips generated or produced in zone i and Dj are the trips attracted
P
P
by zone j. Note that i Oi j Dj . The two constraints in (A.4) indicate that trip generation and attraction for each zone is a
constant, and thus ensure the continuity of trip ows in the system.
If we apply a monotone increasing transformation to the expression for N, more specically by taking the natural logarithm, we obtain

ln N ln

T!

Pi Pi T ij !

ln T! 

XX
i

ln T ij !

A:5

Using Stirlings approximation for the factorial, we have

Nn  en n! ! ln n!  nln n  1

A:6

Finally, given that T is constant, the most likely distribution is obtained by solving the following simple optimization
problem:

max

F

s:t:

fT ij g

T ij ln T ij  1
li

T ij Oi
T ij Dj

F

fT ij g

ij

min

! s:t:

T ij ln T ij  1

ij

T ij Oi

li

T ij Dj

kj

kj

A:7

Interestingly, if we ignore the origin and destination constraints in (A.7), the model will distribute trips uniformly among all
of the origindestination pairs (i, j).
If, on the other hand, we add another objective to minimize total system costs, we arrive at the problem that gives rise to
the classic gravity trip distribution model (Wilson, 1970):

F1

min

F2

s:t:

min
fT ij g

C ij T ij

ij

fT ij g

T ij ln T ij  1

fT ij g

ij

T ij Oi

min

T ij Dj

li

! s:t:

kj

Z1
P

P
ij

C ij T ij 1b

T ij Oi

li

T ij Dj

kj

T ij ln T ij  1

ij

A:8

T ij 0 8i; j

T ij 0 8i; j

By incorporating additional terms into the objective function (A.8) we obtain the other models formulated in this article.
Appendix B. Distribution model based on minimization of the sum of squared differences
The general formulation of this model (Morrison and Thumann, 1980) is based on minimizing the sum of the squared
differences between modeled trips (Tij) and observed trips (Nij) subject to ow conservation constraints for the various zones.

min Z 2
fT ij g

s:t:

1 X T ij  Nij 2
2 ij
pij

A:9

T ij Oi

li

A:10

T ij Dj

kj

A:11

T ij fk

ak

A:12

X
i

X
ij2U k

T ij 0 8i; j wij

A:13

The equation ij T ij fk is an optional added constraint. This restriction gives more degrees of freedom to the model (more
parameters), and it allows to add specic demand structures into the problem, making it possible to capture specic characteristics of the travel pattern, i.e. intra-zone trips, external trips, etc. The values for Nij can be obtained from a survey or an a
priori trip matrix. The parameters pij are weights dened by the modeler that specify the relative importance of each observation. Typical examples of weights are

pij Nij ; N2ij ; Nij ; 1=C ij

The solution to problem (A.9)(A.13) is

A:14

L. de Grange et al. / Transportation Research Part E 46 (2010) 6175

T ij Nij li cj pij

75

ak dij;k pij wij pij

A:15

where dij,k is 1 if Tij falls within the restraint k, and 0 otherwise. Thus, given the observed values (Nij), the weights (pij) and the
Lagrange multipliers (ak, li, cj and wij), we obtain the modeled values Tij. Observe that if the origin-destination (OD) pairs
(i, j) are numerous, many parameters will have to be estimated. Also, the non-negative ow constraint T ij 0 is required to
ensure (A.15) does not generate negative values for Tij.
It should also be noted that this model does not consider the cost of traveling between the various pairs (i, j) in the derivation of the trip distribution. Rather, it is aimed at reproducing as faithfully as possible the observed trip structure of a
given trip matrix, whether of the a priori or expanded survey type. Nevertheless, as we shall see below, objective function
(A.9) is closely related to other distribution models.
References
Abrahamsson, T., Lundqvist, L., 1999. Formulation and estimation of combined network equilibrium models with applications to Stockholm. Transportation
Science 33 (1), 80100.
Beckman, M.J., McGuire, C.B., Winsten, C.B., 1956. Studies in the Economics of Transportation. Yale University Press, New Haven, Connecticut.
Cascetta, E., Pagliara, F., Papola, A., 2007. Alternative approaches to trip distribution modelling: a retrospective review and suggestions for combining
different approaches. Papers in Regional Science 86 (4), 597620.
Cohon, J.L., 1978. Multiobjective Programming and Planning. Academic Press, INC, London.
Congdon, P., 1992. Aspects of general linear modeling of migration. The Statistician 41 (2), 133153.
De Cea, J., Fernandez, J.E., De Grange, L., 2008. Combined models with hierarchical demand choices: a multi-objective entropy optimization approach.
Transport Reviews 28 (4), 415438.
Erlander, S., Stewart, N.F., 1990. The Gravity Model in Transportation Analysis. VSP Utrecht.
Evans, S., 1976. Derivation and analysis of some models for combining distribution and assignment. Transportation Research 10 (1), 3757.
Fang, S.C., Tsao, S.J., 1995. Linearly-constrained entropy maximization problem with quadratic cost and its applications to transportation planning
problems. Transportation Science 29 (4), 353365.
Flowerdew, R., Aitken, M., 1982. A method of tting the gravity model based on Poisson distribution. Journal of Regional Science 22 (2), 191202.
Fotheringham, A.S., 1983. A new set of spatial interaction models: the theory of competing destinations. Environment and Planning 15A (1), 1536.
Fotheringham, A.S., 1986. Modeling hierarchical destination choice. Environment and Planning 18A (1), 401418.
Fotheringham, A.S., OKelly, M.E., 1989. Spatial Interaction Models: Formulations and Applications. Kluwer Academic Publishers, Dordrecht.
Gonalves, M., Souza, J.E., 2001. Parameter estimation in a trip distribution model by random perturbation of a descent method. Transportation Research
35B (2), 137161.
Guldmann, J.M., 1999. Competing destinations and intervening opportunities interaction models of inter-city telecommunication ows? Papers in Regional
Science 78 (2), 179194.
Ham, H., Kim, T., Boyce, D., 2005. Implementation and estimation of a combined model of interregional, multimodal commodity shipments and
transportation network ows. Transportation Research 39B (1), 6579.
Hitchcock, F.L., 1941. The distribution of a product from several sources to numerous localities. Journal of Mathematics and Physics 20 (1), 224230.
Hu, P., Pooler, J., 2002. An empirical test of the competing destinations model. Journal of Geographical Systems 4 (3), 301323.
Knudsen, C.D., Fotheringham, A.S., 1986. Matrix comparison, goodness-of-t and spatial interaction modeling. International Regional Science Review 10 (2),
127147.
Morrison, W.I., Thumann, R.G., 1980. Lagrangian multiplier approach to the solution of a special constrained matrix problem. Journal of Regional Science 20
(3), 279292.
Ortzar, J. de D., Willumsen, L.G., 2001. Modelling Transport. John Wiley & Sons, Chichester.
Roy, J.R., Thill, J.C., 2004. Spatial interaction modelling. Papers in Regional Science 83 (1), 339361.
Schneider, M., 1959. Gravity models and trip distribution theory. Papers, Regional Science Association, 5 (1), 5156.
SECTRA, 2001. Actualizacin de Encuestas de Origen y Destino de Viajes, V Etapa. In: Study conducted by the Direccin de Investigaciones Cientcas y
Tecnolgicas (DICTUC) of the Ponticia Universidad Catlica de Chile.
Sen, A., 1986. Maximum likelihood estimation of gravity model parameters. Journal of Regional Science 26 (3), 461474.
Sen, A., Matuszewski, Z., 1991. Maximum likelihood estimates of gravity model parameters. Journal of Regional Science 31 (4), 469486.
Smith, T.E., 1990. Most-probable-state-analysis: a method for testing probabilistic theories of population behaviour. In: Chatter, J.I.M., Kuenne, R.E. (Eds.),
New Frontiers in Regional Science. MacMillan, London.
Stouffer, S.A., 1940. Intervening opportunities: a theory relating mobility and distance. American Sociological Review 5 (6), 845867.
Thorsen, I., Gitlesen, J.P., 1998. Empirical evaluation of alternative model specications to predict commuting ows. Journal of Regional Science 38 (2), 273
292.
Van Zuylen, J.H., Willumsen, L.G., 1980. The most likely trip matrix estimated from trafc counts. Transportation Research 14B (3), 281293.
Willumsem, L.G., 1981. Simplied transport models based on trafc counts. Transportation 10 (3), 257278.
Willumsen, L.G., 1978. Estimation of an OD matrix from trafc counts: a review. Institute for Transport Studies, Working paper no. 99, Leeds University.
Willumsen L.G., 1984. Estimating time-dependent trip matrices from trafc counts. In: Ninth International Symposium on Transportation and Trafc
Theory. VNU Science Press, Utrecht, pp. 397411.
Wilson, A.G., 1970. Entropy in Urban and Regional Modeling. Pion, London.
Winkelmann, R., Zimmermann, K.F., 1995. Recent developments in count data modeling: theory and application. Journal of Economic Surveys 9 (1), 124.

Das könnte Ihnen auch gefallen