ArticlePDF Available

Comparison of Vehicle-Ownership Models

Authors:

Abstract

Empirical studies on household car ownership have used two types of discrete choice modeling structures: ordered and unordered. In ordered response structures, such as the ordered logit and ordered probit models, the choice of the number of household vehicles arises from a unidimensional latent variable that reflects the propensity of a household to own vehicles. Unordered response structures are based on the random utility maximization principle, which assumes a household associates a utility value across different car ownership levels and chooses the one with the maximum utility. The most common unordered response models are the multinomial logit and probit models, but only the multinomial logit has been used in practical applications because of its simple structure and low computational requirements. Consensus among researchers on unordered or ordered structures is still lacking. Empirical studies have reported various models, including the multinomial logit, ordered logit, and ordered probit. An open question remains: Which model would better reflect households' car ownership choices? This paper compares multinomial logit, ordered logit, and ordered probit car ownership models through a number of formal evaluation measures and empirical analysis of three data sets: the 2001 National Household Travel Survey for the Baltimore [Maryland] Metropolitan Area, the 2005 Dutch National Travel Survey, and the 2000 Osaka [Japan] Metropolitan Person Trip Data. Results show the multinomial logit model should be selected for modeling the level of household car ownership.
THIS IS A COPY OF THE EARLIER VERSION OF THE MANUSCRIPT. THE
PUBLISHED VERSION:
Potoglou, D. and Susilo, Y.O. (2008) Comparison of vehicle-ownership models.
Transportation Research Record, No. 2076, pp. 97 105.
Comparison of Vehicle-Ownership Models
Dimitris Potoglou, Ph.D.
(Corresponding Author)
Delft University of Technology
OTB Research Institute for Housing, Urban and Mobility Studies
P.O. Box 5030
2600 GA Delft
The Netherlands
Tel. +31 15 2785833
Fax. +31 15 2784422
E-mail: d.potoglou@tudelft.nl
Yusak O. Susilo, Ph.D.
Delft University of Technology
OTB Research Institute for Housing, Urban and Mobility Studies
P.O. Box 5030
2600 GA Delft
The Netherlands
Tel. +31 15 2782714
E-mail: y.susilo@tudelft.nl
Word Count: Text (5,586) + 5 Tables (1,250) = 6,836
Paper submitted for presentation and publication at the 87th Annual Meeting of the
Transportation Research Board, Washington, D.C., January 2008
July 2007
Potoglou and Susilo
2
Abstract
Empirical studies on household car ownership have used two types of discrete choice
modelling structures, the ordered and the unordered. In ordered structures such as the ordered
logit and ordered probit models, the choice of the number of household-vehicles arises from a
uni-dimensional latent index that reflects the propensity of a household to own vehicles.
Unordered response models, on the other hand, are based on the random utility maximization
principle, under which a household associates a utility value across different car ownership
levels and chooses the one with the maximum utility. The multinomial logit and probit
models are representatives of the unordered response models, but only the multinomial logit
has been used extensively because of its simple structure and low computational
requirements. Following Bhat and Pulugurta's (1998) comparative study between the
multinomial logit and ordered logit models, consensus among researchers is still lacking and
empirical studies have reported car ownership model based on the multinomial logit, ordered
logit as well as ordered probit models. It is apparent that there is still an open question to be
addressed: Which of the aforementioned models would reflect better households' car
ownership choices? This paper provides an empirical comparison of multinomial logit,
ordered logit and ordered probit car ownership models by introducing a number of formal
evaluation measures and using three datasets; the 2001 National Household Travel Survey for
the metropolitan area of Baltimore, the 2005 Dutch National Travel Survey and the 2000
Osaka Metropolitan Person Trip Data. Results show that the multinomial logit model is the
one to be selected for modelling the level of household car ownership over ordered logit and
ordered probit.
Keywords: car ownership, discrete choice models, automobile demand, multinomial logit,
ordered logit, ordered probit
1. INTRODUCTION
Car ownership is a key feature of modern life that to a great extent influences travel
behaviour and participation in out-of-home activities (1). Increasing utilization patterns and
high-levels of household car ownership have a direct effect on energy consumption and air
quality levels at local and global scales (2). It is therefore of interest to transport planners and
policy makers at all levels of governance to use models that are capable of explaining the
causal factors of car ownership, being sensitive to policy measures and forecast market shares
of automobiles. Such models can be used as stand-alone systems or as part of integrated land-
use and transportation models for simulating interactions between land-use and transportation
(3). Models of car ownership analysis can be distinguished into two broad categories:
disaggregate and aggregate. Disaggregate models focus on the household level, whereas
aggregate models consider car ownership as an accumulation of household decisions at
different geographical scales such the traffic analysis zone, region, state or country levels.
Disaggregated models have dominated over aggregate models mainly because of their
behavioural structure and improved capability in identifying causal relationships (4).
Moreover, disaggregate models have overcome deficiencies and limitations of aggregate
models such as mutli-collinearity across explanatory variables, large standard errors of
Potoglou and Susilo
3
estimated parameters and aggregation bias. Furthermore, the extensive use of disaggregate
models lies in their capability to conduct policy-sensitive analyses and compatibility with
contemporary agent-based approaches in transportation modelling (5).
Disaggregate models regard car ownership observations (i.e., 0, 1, 2, 3 or more cars)
either as ordinal or nominal discrete variables, thus giving rise to two types of choice models,
the ordered and unordered, respectively. Ordered response models assume that the choice of
the number of household-vehicles arises from a uni-dimensional latent index reflecting the
propensity of a household to own vehicles. On the other hand, unordered response models
follow the random utility maximization axiom, in which a household associates a utility value
across different ownership levels and chooses the one with the maximum utility.
A review of the recent literature on disaggregate car ownership models reveals that it
is not yet clear whether ordered response or unordered response models are the most
appropriate (see for example, 6). While Bhat and Pulugurta (7) argued that a multinomial
logit model (MNL) model would be more appropriate for modelling car ownership over an
ordered logit (ORL) model, subsequent empirical studies have developed ordered probit
(ORP) or ORL models claiming the discrete, ordered nature of the dependent variable (8,9) as
well as MNL models on the basis of their sound behavioural and theoretical nature (10-12).
In particular, comparing modelling results of MNL and ORL using several datasets, Bhat and
Pulugurta (7) found substantial differences in the elasticities of exogenous variables across
the choice probabilities of car ownership levels and identified misspecification problems
associated with the ORL that could lead to incorrect and inaccurate forecasts. Also, Potoglou
and Kanaroglou (12) found that a MNL model was a significantly improved model over the
ORL through a likelihood-ratio test between the two models using data from the metropolitan
area of Hamilton, Canada.
This paper offers a comprehensive comparison of car ownership models including
MNL, ORL and ORP. Two research questions are addressed in the paper. First, what
measures could be used in order to conduct a comparison between different discrete choice
model structures? And second, which is the most appropriate among the MNL, ORL and
ORP for modelling car ownership? We address these questions by developing a set of the
aforementioned car ownership models using three sources of data; 2001 U.S. National
Household Travel Survey (NHTS) of the Baltimore Metropolitan Area (13), the 2005 Dutch
National Travel Survey (14) and the 2000 Osaka Metropolitan Person Trip Data (15).
Following, we compute a number of comparison measures introduced in the econometrics
literature.
The reminder of the paper is organized as follows. First, we provide a brief
description on the structure and major assumptions of ordered and unordered models. Next,
we introduce a set of evaluation measures to be used in the comparison of models. Following
that, we present the data sets used and describe the construction of explanatory variables.
Next, we discuss estimation results of each model and report on the findings of the
comparative analysis as reflected by the evaluation measures. In the last section, we offer
some concluding remarks.
2. ORDERED AND UNORDERED MODELS: THEORETICAL BACKGROUND
Ordered and unordered models require different techniques for their respective analysis (16).
We briefly present the main characteristics and assumptions of ordered and unordered models
in the following subsections.
Potoglou and Susilo
4
2.1 Ordered Response Models: ORL and ORP
Ordered response models assume that the observed number of household cars (dependent
variable) is a discrete, ordinal variable that is mutually exclusive and collectively exhaustive.
Hence, the observed number of cars per household is assumed to be inherently ordered
implying that higher number of cars is ranked higher than the outcome associated with less
number of cars. Specifically, the ordered response models assume that the number of vehicles
of household n - denoted as (Yn) - arises from a dimensional latent index
*
n
y
, as follows :
n
K
1n nkk
*
nXy
[1]
where,
*
n
y
reflects the propensity of a household to own vehicles, k (to be estimated) is a set
of parameters associated with a set of k explanatory variables (Xnk). The distribution of the
error terms
n
marks the difference between the ordered logit and the ordered probit models.
In general, the observed level of automobile ownership of a household (Yn) equal to i
( = 0, 1,…M) number of cars is assumed to be related to the latent auto ownership index as
follows (17):
iallfor,,0,,M,...,1,0i,yifonlyandifiY i1i0i1i
*
n1in 
[2]
where are cut-offs points for discriminating between successful automobile ownership
levels on the underlying latent scale. Cut-off points are estimated along with the
coefficients in one-step procedure (see, 18). Hence, the probability of household n to own i
number of cars is as follows:
)x()x(P]iY[P n1ininin
[3]
where (.) is the standard cumulative normal distribution, which gives rise to the ordered
probit model. Alternatively, (.) can be replaced with the Gumbel distribution, (.),
resulting into the ordered logit model. With regard to the selection of the cumulative
distribution function and hence, the choice between ordered logit and probit models, Greene
(18) pointed out that "it is difficult to justify the choice of one distribution over the other on
theoretical grounds…in most applications, it seems not to make much difference".
A significant assumption in the estimation of ordered response models (ordered probit
and logit) is that of parallel slopes. This assumption implies that the estimated coefficient of
an explanatory variable affecting the probability of a household to own a number of cars
would be equal for all outcomes (i.e., i = 1, 2, 3 or more cars) (16). If the parallel slopes
assumption is invalid and coefficients associated with a particular variable are different
across different levels of car ownership, then the ordered response mechanism is no longer
appropriate. In that case, the model should be estimated using an unordered response model.
As Borooah (16) suggested "the validity of the parallel slopes assumption can be tested by
estimating a multinomial logit model on the data…while the ORL model estimates K
coefficients, the MNL model estimates K(M-1) parameters…if the LL1 is the likelihood
function value from the ORL model and LL2 is the likelihood value from the MNL model,
then one can compute 2*(LL2-LL1) and compare with 2(K(M-2))". This test is only
suggestive implying that a "very large" 2 value would provide grounds for concern, a
moderately "large" value would not (16).
Potoglou and Susilo
5
2.2 Unordered Response Models: MNL and MNP
In the case of the unordered response models, the information conveyed by the ordinal nature
of the observed car household ownership is discarded. Hence, the probability of a household
to own a given number of automobiles is based on the utility maximization principle, in
which a household n associates a utility value to each car ownership level and chooses the
one with the maximum utility. Assuming that a household is perfectly informed and the
decision is completely rational and consistent, the random utility maximization framework
implies that:
n,inn,in,i XU
[4]
where Ui,n represents the true utility of a household n owning i number of cars, i,n is a vector
of parameters to be estimated and i,n is the random component that captures the unobserved
utility or uncertainty from the point of view of the observer/analyst. Again, if the error terms
i,n in Equation 4 are identically and independently distributed (IID) with a Type I Extreme
Value (or Gumbel) distribution, the probability of owning a number of cars i takes the form
the MNL expressed as follows (19):
n
Cj nn,j
nn,i
n)xexp(
)xexp(
)i(P
[5]
The computational simplicity of the MNL model has permitted its wide use in empirical
studies of car ownership. Nevertheless, the MNL has an undesirable property known as the
"independence of irrelevant alternatives" (IIA). The IIA property implies that the ratio of the
probabilities of two alternatives is independent from any other available alternatives. Should
this assumption is violated then the MNL is unsuitable for use in any application in which the
random components of utility are correlated across alternatives and observations of choices
(20). Alternatively, the assumption of error terms i,n being multivariate normal distributed
with mean zero and covariance matrix n leads to the Multinomial Probit (MNP) model. As
opposed to the MNL, the MNP model is more flexible because it relaxes the IIA assumption
of IID errors and thus, allows for correlations of the error terms of different alternatives (20).
In the MNP model, the probability of a household choosing a certain number of cars i is
given as follows:
[6]
where
n,i
~
is a vector of error differences
nJii1nn,i ~
,...,
~~
over all alternatives except i,
with density function
n,i
~
.
As shown in Equation 6, the required integrations are computationally burdensome,
especially when the number of alternatives increases. As a result, the use of the MNP model
has been limited in empirical studies involving discrete choice models. Several simulation
methods (see, 21) have been introduced to overcome this problem, however, there remain
considerations with regard to the applicability of the MNP model. Weeks (22) concluded that
the estimation of the MNP entails several estimation, specification and identification issues.
Also, Horowitz (23) argued that the MNP model involves problems such as the proliferation
Potoglou and Susilo
6
of random effects and parameters as well as specification testing. Proliferation can
complicate forecasting as well as adds to the complexity of estimation because it requires a
higher number of covariance matrix elements to be estimated. Regarding specification
testing, Horowitz (23) claimed that "… a misspecified model can create non-IID random taste
variation or additive random components in the error terms of the model". In such case, it
may appear that a MNP model with random taste variation or non-IID additive random utility
components fits the data set better that a MNL model, when the real problem is that the
systematic component of the logit utility function is misspecified. As a result, the use of MNP
for car ownership models may present several difficulties, while also estimation results may
not be directly comparable with those of a MNL.
TABLE 1 Main characteristics of common ordered and unordered models
Ordered Models
Unordered Models
ORL
ORP
MNL
MNP
Behavioural Context
No
No
Yes
Yes
Distribution of Error Terms
Gumbel
Normal
IID - Gumbel
Normal
Parallel Slopes Assumption
Yes
Yes
No
No
IIA Property
No
No
Yes
No
Computational Requirements
Low
Low
Low
High
3. EVALUATION MEASURES OF DISCRETE CHOICE MODELS
The likelihood ratio
2
and the adjusted likelihood ratio indices
2
are the most common
measures of goodness-of-fit used with discrete choice models and are given as follows (20):
)C(LL
)
ˆ
(LL
1
2
[7]
)C(LL
K)
ˆ
(LL
1
2
[8]
where
)
ˆ
(LL
and
)C(LL
are the log-likelihood function values at convergence and sample
shares, respectively. K is the number of parameters estimated in the model, excluding any
constants as well as cut-off parameter values in the case of ordered models.
The difference [
K)
ˆ
(LL
] is also known as the Akaike Information Criterion (AIC),
which is another measure of goodness-of-fit of a model given as (24):
K2)
ˆ
(LL2AIC
[9]
To correct for small samples Hurvich and Tsai (25) proposed the AICc criterion, which can
be used regardless of sample size:
)1KN(
)1K(K2
K2)
ˆ
(LL2AICc
[10]
Potoglou and Susilo
7
where N is number of observations. AIC and AICc allow for a better comparison across
models because they account for goodness-of-fit as well as include a penalty that is an
increasing function of the number of parameters.
Another measure for model comparison is the Bayesian Information Criterion (BIC),
also known as Schwarz Criterion (SIC) (26):
)Nln(K)
ˆ
(LL2BIC
[11]
The BIC can be used to compare models that are non-nested as is the case with MNL, ORL
and ORP models. Given any estimated models derived from the same sample, the model with
the lower value of AICc or BIC is the one to be preferred because this model is estimated to
be the "closest" to the unknown true model. Compared with AIC, the BIC penalizes free
parameters more strongly than the AIC.
Another consistent criterion as opposed to the non-consistent AICc and BIC is the
Hannan and Quinn Information Criterion (HQIC) criterion (27):
))Nln(ln(K2)
ˆ
(LL2HQIC
[12]
AICc, BIC and HQIC indices take into account model parsimony. That is, other things equal,
given two models with equal log-likelihood values, the model with fewer parameters is
better. Finally, a comparison between two non-nested models may be performed with Ben-
Akiva and Lerman's (20) adjusted log-likelihood ratio test, which determines if the adjusted
likelihood ratio indices between two non-nested models are significantly different. Under the
null hypothesis that model 1 is a better representation than model 2, the following holds
asymptotically:
5.0
12
2
1
2
2)KK()C(LLz2)zPr(
, z > 0 [13]
where (.) is the standard normal cumulative distribution function. As Bhat and Pulugurta
(7) argued a small value of the probability in Equation 13 indicates that the difference z is
statistically significant and the model with the higher value of adjusted likelihood ratio index
is to be preferred.
Finally, another goodness-of-fit statistic is "the percent correctly predicted", which is
calculated by identifying the highest probability for each decision-maker in the sample. The
"the percent correctly predicted" is the portion of decision-makers in the sample for which the
alternative with the highest probability and the actually chosen alternative are the same.
However, as Train (21) argued, this contradicts with the notion of probability. A model
provides enough information in identifying the probability of choosing each alternative,
however, it is incapable of determining repeatedly the actual choice. Hence, this measure will
not be considered further in the evaluation of models.
4. SOURCES OF DATA
The data sets used in this paper were obtained from three sources: the 2001 US NHTS (13),
the 2005 Dutch NTS (14) and the 2000 Osaka Metropolitan Person Trip Data (15). The US
NHTS has been conducted by the US Bureau of Transport Statistics (BTS) and the US
Federal Highway Administration (FHWA). The objective of this survey is to gain a better
understanding of travel behaviour in the US. The data are used by the US Department of
Transportation to assess program initiatives, to review programs and policies, to study current
mobility issues, and to plan for the future. In this paper, the analyses used data from the
Potoglou and Susilo
8
Baltimore Metropolitan Area, which consist of 3,496 households, 8,744 persons, and 32,359
trips. The Dutch NTS is a cross-sectional travel-diary of households and data collection has
been conducted continuously by Statistics Netherlands since 1978. For each year up to 1993,
the NTS recorded data for approximately 10,000 households, 20,000 individuals (and more
than 80,000 journeys). During 1994 and 1995 the NTS was extended to include substantially
more respondents and households each year and also to include children younger than 12,
who were previously excluded from the survey (14).
Finally, the Osaka metropolitan area person-trip survey is a conventional large-scale
household travel survey conducted in the Osaka metropolitan area of Japan in 1980, 1990 and
2000, with sampling rates of 2.4% to 3.0%. The survey collected travel patterns as well as
home and work locations of the respondents on the observed day. It also collected the socio-
demographics and household characteristics (15).
5. ESTIMATION RESULTS AND EVALUATION OF MODELS
Car ownership models were estimated considering four car ownership levels: zero, one, two
and three or more cars. All estimations were performed using the program NLOGIT Ver. 3.0
(28) and final models are reported in Tables 2, 3 and 4. There are some key features to be
addressed at this point. First, parameters of the MNL models were estimated for three
alternatives (i.e., one, two and three or more cars), whereas the zero-cars alternative was
considered as base alternative for identification purposes. Second, the parameters of the MNL
models indicate propensity to own one, two or three or more vehicles in such way that
positive values increase the probability and negative values decrease the probability of
owning a particular number of cars. Similarly, the coefficients in the ordered response models
indicate propensity to own more cars and thus, positive values of coefficients indicate
propensity of households to own higher number of cars. Finally, it is worth noting that MNP
models of car ownership did not reach convergence with any of the aforementioned datasets,
and therefore MNP model estimated are not reported in this comparison.
With regard to the Baltimore dataset, the single-family house attribute was used as
proxy for parking-space availability at the place of residence. The coefficient of this variable
was significant in all models, except in the utility of owning one car. Thus, in the case of
MNL model, households living in single family dwellings were more likely to own two and
three-or-more cars. In the ORL and ORP models, the positive sign of the Single Family
House attribute implied the increased propensity of households to own cars. This finding is in
line with findings reported in Potoglou and Kanaroglou (12), Bhat and Pulugurta (7) and Chu
(8). Furthermore, socio-economic variables such as number of workers, total household
income and race of the head of the household had positive coefficients implying an increased
probability of owning vehicles in the MNL as well as an increased likelihood of owning more
vehicles in the case of ORL and ORP models, respectively. Furthermore, as shown in Table
3, parameter signs of the number of workers and income variables were consistent with
estimation results using the 2005 Dutch NTS.
The household life cycle reflected the influence of the composition of the household
on the number of cars owned. In all datasets, we used as reference class the household
composition "single" (households with one adult only). As shown in Table 2, the MNL model
captures non-linear effects between household life cycle and the probabilities of car
ownership. Specifically, a household comprised of two adult members ("couple") is less
Potoglou and Susilo
9
likely to own one car and more likely to own two or three-or-more cars. It is demonstrated in
this case that unordered models such as the MNL placed no restrictions on the effect of an
explanatory variable across car ownership levels. This finding agrees with the theoretical
claim of Bhat and Pulugurta (7).
Also in Table 2, the pattern of switching parameter signs of the variable "couple"
across car ownership levels cannot be captured in ORL and ORP models where the parameter
value is restricted to a unique (positive) parameter. Similarly, the "couple with children"
variable had no effect in the probability of owning one car, however, it did affect positively
the probability of owning two-or-more vehicles. Furthermore, the single parent household
structure had no effect at all levels of car ownership; that was also the case of "retired"
households and the probability of owning one car. For all other types of household structure,
coefficients were statistically significant and their signs agreed across MNL, ORL and ORP
models as well as with a priori expectations.
TABLE 2 Parameter estimates of the Baltimore Dataset
MNL
ORL
ORP
Variable
1
2
3 +
Constant
0.504
(1.94)
-2.695
(-8.00)
-6.746
(-12.52)
-0.374
(-2.32)
-0.191
(-2.09)
Type of Dwelling: Single Family House
0.275
(1.49)
1.118
(5.53)
1.778
(7.38)
0.763
(8.36)
0.452
(8.77)
Number of Workers
0.509
(4.13)
1.212
(8.65)
2.175
(13.77)
1.035
(17.15)
0.574
(17.15)
Household Income (>$30,000)
REFERENCE
Household Income ($30,000 - $80,000)
1.683
(10.80)
2.401
(13.18)
3.178
(11.16)
1.413
(15.44)
0.806
(15.82)
Household Income (> $80,000)
2.013
(4.91)
3.708
(8.80)
4.659
(9.82)
1.906
(16.61)
1.100
(16.91)
Household Life Cycle: Single
REFERENCE
Household Life Cycle: Couple
-0.573
(-2.46)
1.872
(6.91)
1.950
(4.73)
1.243
(10.41)
0.669
(9.84)
Household Life Cycle: Single Parent
-0.333
(-1.49)
-0.070
(-0.21)
-0.254
(-0.42)
-0.080
(-0.48)
- 0.053
(-0.55)
Household Life Cycle: Couple w. Children
-0.460
(-1.81)
2.082
(7.14)
2.350
(5.51)
1.383
(10.97)
0.751
(10.49)
Household Life Cycle: Retired
0.011
(0.06)
1.864
(8.07)
2.620
(6.73)
1.310
(11.83)
0.733
(11.60)
Race: Caucasian
1.066
(7.95)
1.636
(9.45)
2.007
(8.45)
0.939
(10.53)
0.531
(10.56)
Residential Density
-0.00025
(-6.62)
-0.00046
(-10.85)
-0.00063
(-11.63)
-0.00025
(-12.82)
-0.00015
(-13.02)
Threshold Parameters
(for identification 0=0)
1
3.087
(52.76)
1.726
(53.49)
2
6.263
(74.16)
3.520
(79.60)
Log-likelihood at convergence
-2949
- 3081
-3092
Potoglou and Susilo
10
TABLE 3 Parameter estimates using the 2005 Dutch NTS Dataset
MNL
ORL
ORP
Variable
1
2
3 +
Constant
-1.051
(-20.24)
-5.649
(-47.67)
-9.090
(-23.50)
-1.065
(-24.36)
-0.590
(-23.20)
Number of Workers
0.964
(22.09)
1.559
(31.19)
2.120
(27.62)
0.888
(38.74)
0.480
(37.52)
Household Income (>€20,000)
REFERENCE
Household Income (€20,000 - €60,000)
0.800
(21.72)
1.593
(22.88)
1.312
(5.30)
0.885
(27.38)
0.500
(27.47)
Household Income (>€60,000)
1.693
(15.76)
3.269
(26.61)
3.475
(12.94)
1.791
(39.26)
1.020
(39.48)
Household Life Cycle: Single
REFERENCE
Household Life Cycle: Couple
1.073
(20.83)
2.757
(28.14)
1.534
(5.27)
1.256
(31.29)
0.700
(31.06)
Household Life Cycle: Couple w. Children
1.425
(18.04)
3.596
(31.64)
1.904
(6.23)
1.722
(39.12)
0.951
(38.64)
Household Life Cycle: Extended family
0.687
(8.95)
2.779
(24.29)
3.298
(11.60)
1.856
(36.93)
1.030
(37.17)
Household Life Cycle: Retired
0.137
(2.88)
0.577
(4.71)
0.416
(0.97)
0.159
(3.78)
0.09
(3.52)
Reside in very highly urbanised area
REFERENCE
Reside in high urbanised area
0.631
(12.36)
1.026
(12.95)
1.550
(6.35)
0.616
(14.79)
0.350
(14.91)
Reside in moderately urbanised area
0.904
(16.25)
1.557
(18.85)
2.199
(9.01)
0.911
(21.01)
0.520
(21.23)
Reside in low urbanised area
1.195
(20.63)
1.987
(23.73)
3.049
(12.81)
1.168
(26.933)
0.670
(27.22)
Reside in non-urbanised area
1.306
(20.29)
2.258
(25.09)
3.209
(13.07)
1.301
(28.17)
0.740
(28.36)
Threshold Parameters
(for identification 0=0)
1
3.870
(153.13)
2.19
(169.49)
2
7.145
(162.78)
3.95
(185.95)
Log-likelihood at convergence
-22941
-23255
-23254
Potoglou and Susilo
11
TABLE 4 Parameter estimates using the 2000 Osaka Metropolitan
Area Dataset
MNL
ORL
ORP
Variable
1
2
3 +
Constant
-0.397
(-14.53)
-1.541
(-35.89)
-2.318
(-35.08)
0.520
(21.77)
0.325
(23.93)
Household Life Cycle: Single
REFERENCE
Household Life Cycle: Couple
0.912
(29.59)
0.252
(6.77)
-1.235
(-18.74)
-0.356
(-18.59)
-0.211
(-18.30)
Household Life Cycle: Single parent
-0.871
(-19.10)
-2.843
(-25.00)
-4.388
(-16.21)
-1.953
(-45.65)
-1.154
(-45.02)
Household Life Cycle: Couple w.
Children
0.687
(34.88)
0.011
(0.46)
-1.373
(-34.62)
-0.465
(-35.52)
-0.272
(-35.24)
Household Life Cycle: Extended family
1.164
(81.37)
1.092
(69.87)
0.835
(48.35)
0.299
(36.01)
0.190
(39.82)
Household Life Cycle: Retired
-1.061
(-85.54)
-2.195
(-118.0)
-3.305
(-105.86)
-1.898
(-184.89)
-1.088
(-186.24)
Reside in CBD area
REFERENCE
Reside in mixed area
0.564
(20.30)
0.577
(13.09)
0.586
(8.58)
0.484
(19.85)
0.267
(19.27)
Reside in autonomous area
1.458
(41.19)
3.212
(66.16)
4.200
(59.87)
2.631
(101.21)
1.499
(100.93)
Reside in suburban area
1.270
(46.46)
2.011
(46.72)
2.362
(35.61)
1.452
(60.97)
0.824
(60.95)
Reside in non-urbanised area
1.792
(22.75)
3.398
(39.19)
4.624
(45.99)
2.804
(69.53)
1.609
(68.43)
Threshold Parameters
(for identification 0=0)
1
2.375
(531.86)
1.400
(555.86)
2
3.877
(659.11)
2.265
(710.88)
Log-likelihood at convergence
-350228
-355634
-355820
Finally, we specified a full set of alternative specific constants in the MNL models
corresponding to one, two and three-or-more-cars options in order to capture the systematic
influence of omitted variables, namely car ownership costs, in the utility functions of the
model (8).
5.1 Behavioural Comparisons across the American, Dutch and Japanese Datasets
Comparing the estimated coefficients across the Baltimore, Dutch and Japanese
datasets, it is clear that estimates present some discrepancies, which may be explained by the
contextual/cultural differences in the sampled population and also, by the different
explanatory variables available in the dataset. For example, while in the Baltimore dataset
(Table 2) we used residential density to measure the influence of land-use on car ownership,
we used the degree of urbanisation in the Dutch (Table 3) dataset and urban-functions in the
dataset of Osaka metropolitan area (Table 4). Moreover, unlike the other two datasets, the
Osaka metropolitan area dataset did not include household income and information on the
number of workers, which is a crucial variable for car ownership decisions.
Potoglou and Susilo
12
Nevertheless, comparison of the datasets reveals interesting differences. For example,
except for the extended family case, the Dutch and Osaka metropolitan respondents tend to
have two cars rather than three or more cars, which is different than Baltimore respondents.
Presumably this is because of the different needs for car usage. Dutch people tend to walk
and to bike in their daily travel (29), while the Osaka metropolitan area has a very dense and
well developed public transport network (15) compared to Baltimore area. Yet, despite the
differences in describing the land-use influences to the household car ownership, results in all
datasets are in line with the hypothesis that higher densities or proximity to the activity
locations reduces the probability of a household owning a car.
Furthermore, examining the influence of household structure to household car
ownership, we found that - all else being equal - while Baltimore couples were less likely to
own one car, Dutch and Osaka's couples were more likely to own one and three cars than
zero cars. Also, single-parent households residing in Baltimore and Osaka had a higher
probability of owning one car and lower probability of owning three cars. On the other hand,
Dutch single parents were more likely to own one car and less likely to posses three cars. As
shown in Tables 2, 3 and 4, couples with children also present differences across the dataset
estimates. In particular, couples with children living in Baltimore were less likely to own one
car and more likely to own three cars, as opposed to Dutch and Osaka's households that
would be more likely to own one car. Interestingly, Baltimore's retirees had a higher
probability to own cars as opposed to Osaka's where the negative coefficients imply that the
probability of owning any number of cars would be less than owning no cars. Again, these
discrepancies are because of differences in public transport availability and high usage of
non-motorized modes across Baltimore, the Netherlands and Osaka. It is worth mentioning
here that these discrepancies are also reflected in the estimates of the ORL and ORP models.
5.2 Comparison of MNL, ORL and ORP Models
All models performed well as indicated by the relatively high values of rho-squared. Also,
likelihood ratio tests were used to test the null hypothesis that all parameters in each model -
except the alternative specific constants - were zero. As shown in Table 5, chi-square values
of the likelihood ratio indices in all models reject the null hypothesis and indicate that all
models were statistically significant. The adjusted likelihood values lend support to the MNL
model as the best for modelling car ownership because it exhibits the highest value of the
index. Also, the values of AICc, BIC and HQIC indices clearly demonstrate the superiority of
the MNL against the ORL and ORP. Finally, Ben-Akiva and Lerman's (20) adjusted log-
likelihood ratio test was used for testing the non-nested hypothesis that the adjusted log-
likelihood values of the MNL, ORL and ORP were statistically different. The last three rows
of Table 5 present the upper bound of the probability of erroneous choosing the incorrect
model, which however, had the highest adjusted likelihood index. In other words, this
hypothesis states that the estimated differences in the adjusted likelihood ratio index values
between the MNL, ORL and ORP models could have occurred by chance. The estimated
probability values of the aforementioned hypothesis showed that the adjusted likelihood ratio
index of the MNL is significantly different than those of the ORL and ORP models and
therefore, the MNL is the preferred model. Similarly, the ORL model should be preferred as
opposed to the ORP model in the Baltimore case, whereas there is no difference between the
ORL and ORP models in the Dutch and Japanese datasets.
Potoglou and Susilo
13
Potoglou and Susilo
14
6. CONCLUSION
Car ownership is a key element in the study and simulation of urban systems. Hence, it is of
interest to develop models of car ownership that are capable of explaining and predicting
households' choices. To date, empirical studies have focused on two types of disaggregate
models, the ordered and unordered. The first type refers to ORL and ORP whereas the second
is mainly represented by the MNL. While a previous comparison suggested the MNL as more
appropriate for car ownership modelling over an ORL (7), subsequent empirical analyses
have adopted both MNL and ORL as well as ORP models. Thus, it has not been clear, which
of the aforementioned models would be more suitable for car ownership modelling.
The objective of this paper has been to evaluate the MNL, ORL and ORP models for
car ownership based on a number of data fit measures. We approached this task by
empirically studying household car ownership levels using as explanatory variables the
household's life-cycle stage, income, race, type of dwelling and number of workers per
household. In addition, we included residential density as a measure of urban form assuming
that higher densities would discourage households to own more vehicles. The results of
model estimations confirm previously published findings highlighting the importance of
socio-demographic and economic characteristics of households as well as the type of
dwelling and urban form in explaining household car ownership.
A key difference between ordered and unordered models is that unordered models,
namely the MNL, are based on the utility maximization principle, whereas ordered models
(ORL, ORP) are not. This difference makes the MNL more appealing over ordered models,
because findings are based on a solid behavioural framework and not a single continuous
propensity measure. Furthermore, a quick glance in the estimation of parameter reveals that
the MNL model is more flexible because it allows for alternative-specific effects of
explanatory variables across car ownership levels. On the other hand, ordered models are
constrained to a unique coefficient per explanatory variable. Finally, comparison tests of data
fit among the MNL, ORL and ORP models suggest that the MNL model should be the
preferred model structure for household car ownership.
ACKNOWLEDGEMENTS
An earlier version of this paper was presented at the 9th NECTAR Conference held in
Porto, Portugal in May 2007. The authors are grateful to Professors Yoram Shiftan of
Technion-Israel Institute of Technology and Gerard de Jong of Institute of Transport Studies
at University of Leeds for their constructive comments and suggestions. Also, the authors
would like to thank Nam Seok Kim, Ph.D. Candidate at the OTB Research Institute of Delft
University of Technology, for his help in identifying the Baltimore data for this analysis and
the Kinki Branch Office of the Japanese Ministry of Land, Infrastructure and Transportation
for providing Osaka's dataset.
REFERENCES
1. Ortuzar, J. D. and L. G. Willumsen Modeling Transport. John Wiley & Sons, Ltd,
Chichester, 2001.
2. Fenger, J. Urban Air Quality. Atmospheric Environment, Vol. 33, 1999, pp. 4877-4900.
Potoglou and Susilo
15
3. Miller, E. J., J. D. Hunt, J. E. Abraham and P. A. Salvini. Microsimulating Urban Systems.
Computers, Environment and Urban Systems, Vol. 28, 2003, pp. 9-44.
4. Bunch, D. S. Automobile Demand and Type Choice In Handbook of Transport Modeling;
Hensher, D. A. and Button, K. J., (Eds.), Pergamon, New York, 2000, pp 463-479.
5. Miller, E. J., D. S. Kriger and J. D. Hunt. Integrated Urban Models for Simulation of
Transit and Land-Use Policies. University of Toronto, Joint Program in Transportation and
DELCAN Corporation, 1998.
6. Potoglou, D. and P. S. Kanaroglou. Disaggregate Demand Analyses for Conventional and
Alternative Fuelled Automobiles: A Review. International Journal of Sustainable
Transportation (forthcoming), 2007
7. Bhat, C. R. and V. Pulugurta. A Comparison of Two Alternative Behavioural Choice
Mechanisms for Household Auto Ownership Decisions. Transportation Research Part B,
Vol. 32, No. 1, 1998, pp. 61-75.
8. Chu, Y.-L. Automobile Ownership Analysis Using Ordered Probit Models. In
Transportation Research Record: Journal of the Transportation Research Board, No. 1805,
TRB, National Research Council, Washington, D.C., 2002, pp. 60-67.
9. Kim, H. S. and E. Kim. Effects of Public Transit on Automobile Ownership and Use in
Households of the USA. RURDS - The Applied Regional Science Conference (ARSC), Vol.
16, No. 3, 2004, pp. 245-262.
10. Ryan, J. M. and G. Han. Vehicle-Ownership Model Using Family Structure and
Accessibility Application to Honolulu, Hawaii. In Transportation Research Record: Journal
of the Transportation Research Board, No. 1676, TRB, National Research Council,
Washington, D.C., 1999, pp. 1-10.
11. Purvis, C. L. Using 1990 Census Use Microdata Sample to Estimate Demographic and
Automobile Ownership Models. In Transportation Research Record: Journal of the
Transportation Research Board, No. 1443, TRB, National Research Council, Washington,
D.C., 1994, pp. 21-29.
12. Potoglou, D. and P. S. Kanaroglou. Modelling Car Ownership in Urban Areas: A Case
Study of Hamilton, Canada. Journal of Transportation Geography (in press/
doi:101016/jjtrangeo200701006), 2007
13. NHTS. National Household Travel Survey: Federal Highway Administration
http://nhts.ornl.gov/download.shtml. Accessed March 2007.
14. van Evert, H., W. Brog and E. Erl. Survey Design: The Past, the Present and the Future In
Travel Survey Methods - Quality and Future Directions; Stopher, P. and Stecher, C., (Eds.),
Elsevier, London, 2006.
Potoglou and Susilo
16
15. Susilo, Y. O. and R. Kitamura. Structural Changes in Commuters' Daily Travel: The Case
of Auto and Transit Commuters in the Osaka Metropolitan Area of Japan, 1980 through
2000. Transportation Research A, 2007 (forthcoming).
16. Borooah, V. K. Logit and Probit. Sage Publications, London, 2001.
17. McKelvey, R. D. and W. Zavoina. A Statistical Model for the Analysis for Ordinal Level
Dependent Variables. Journal of Mathematical Sociology, Vol. 4, 1975, pp. 103-120.
18. Greene, W. H. Econometric Analysis (4th Edition). Prentice Hall, Englewood Cliffs, NJ,
2000.
19. McFadden, D. Conditional Logit Analysis of Qualitative Choice Behaviour In Frontiers
in Econometrics; Zerembka, P., (Ed.), Academic Press, New York, 1973, pp 105-142.
20. Ben-Akiva, M. and S. R. Lerman Discrete Choice Analysis: Theory and Application to
Travel Demand. MIT Press, Cambridge, 1985.
21. Train, K. Discrete Choice with Simulations. Cambridge University Press, Cambridge,
2002.
22. Weeks, M. The Multinomial Probit Model Revisited: A Discussion of Parameter
Estimability, Identification and Specification Testing. Journal of Economic Surveys, Vol. 11,
No. 3, 1997
23. Horowitz, J. L. Reconsidering the Multinomial Probit Model. Transportation Research
Part B, Vol. 25, No. 6, 1980, pp. 433-438.
24. Akaike, H. 2nd International Symposium on Information Theory, Akademiai Kaido,
Budapest, Hungary, 1973, pp 267-281.
25. Hurvich, C. M. and C. Tsai. Regression and Time Series Model Selection in Small
Samples. Biometrika, Vol. 76, No. 2, 1989, pp. 297-307.
26. Schwarz, G. Estimating the Dimension of a Model. Annals of Statistics, Vol. 6, No. 2,
1978, pp. 461-464.
27. Hannan, E. and B. Quinn. The Determination of the Order of an Autoregression. Journal
of the Royal Society Series B (Methodological), Vol. 41, No. 2, 1979, pp. 190-195.
28. Greene, W. H. Nlogit Ver. 3. Econometric Software Inc., New York & Sydney, 2002.
29. Susilo, Y. O. and K. Maat. The Influence of Built Environment to the Trends in
Commuting Journeys in the Netherlands. Transportation (forthcoming), 2007
... Car and vehicle ownership models can be estimated using either an aggregate or disaggregate approach; the household is the basic unit in disaggregate models (70). Disaggregate models are considered more efficient than aggregate models with respect to their structure and improved capability of identifying associations between the number of cars and vehicles per household and convenient explanatory variables (30,70). ...
... Car and vehicle ownership models can be estimated using either an aggregate or disaggregate approach; the household is the basic unit in disaggregate models (70). Disaggregate models are considered more efficient than aggregate models with respect to their structure and improved capability of identifying associations between the number of cars and vehicles per household and convenient explanatory variables (30,70). Those models enable the number of cars and vehicles per household to be represented as either ordinal or nominal discrete variables, such as ordered or unordered response models (70). ...
... Disaggregate models are considered more efficient than aggregate models with respect to their structure and improved capability of identifying associations between the number of cars and vehicles per household and convenient explanatory variables (30,70). Those models enable the number of cars and vehicles per household to be represented as either ordinal or nominal discrete variables, such as ordered or unordered response models (70). Table 4 summarizes selected studies in the existing literature that incorporated a disaggregate modeling approach for their own samples. ...
Article
Car and vehicle ownership levels in developing countries are gradually increasing, leading to traffic congestion, energy consumption, and air pollution. As a developing country, Turkey encounters many of the issues caused by this rapid increase in car ownership. This paper aims to examine potential determinants of household car and vehicle ownership behavior in urban areas of Turkey by concentrating on three cities (Ankara, Erzurum, and Bayburt) that are distinct in population, geographical region, and public transit services. For this purpose, a written survey was administered to 1389 households living in these cities. The data is analyzed using alternative ordered response models—the generalized ordered logit, partial proportional odds (PPO), and heteroskedastic ordered logit models. To the best of our knowledge, this study is the first attempt to compare alternative ordered response models to determine car and vehicle ownership behavior in Turkey. Consequently, the PPO model was found to be the most parsimonious model among the others. The results demonstrated that household car and vehicle ownership behavior in Turkey differs significantly with urban area size. The estimation results also revealed that several key drivers, including type of dwelling, parking availability, household monthly income, and public transit use, significantly influence Turkish households’ car and vehicle ownership levels. This study’s empirical findings may provide insight to policymakers and stakeholders for future transportation strategies and policies.
... Several types of models have been used to understand vehicle ownership in the past, with the multinomial logit (MNL) and ordered logit (OL) or probit (OP) models being the most common structures in the literature (Potoglou and Susilo, 2008). These models use a different discrete choice structure: the MNL has an unordered response structure, and the OL/OP has an ordered response structure. ...
... These models use a different discrete choice structure: the MNL has an unordered response structure, and the OL/OP has an ordered response structure. To define which approximation is better to represent vehicle ownership choice, Bhat and Pulugurta (1998), Potoglou and Susilo (2008) and Iglesias et al. (2022) tested both model forms using different datasets from the United States, Japan, the Netherlands and Colombia. These studies concluded that the MNL model produced the best goodness of fit for each dataset. ...
... We begin with the descriptive statistics and tests of significance to estimate the scope and scale of rural households and residents who have no or limited car access, quantify socioeconomic and mobility differences by household car access, and identify whether disparities in household car access within rural areas differ from that in nonrural areas and to what degree. Building upon prior work, we grouped U.S. households by their degree of car access into three categories: carless, car-deficit, and fully equipped households (Basu and Ferreira 2020;Potoglou andSusilo 2008, Blumenberg et al. 2020). In producing the descriptive summary statistics, we conducted Chi-squared tests of independence for those factors that are categorical variables, such as gender, race and ethnicity, and educational attainment, and used the Kruskal-Wallis rank-sum test for those factors that have a continuous value, such as household income. ...
... We then carried out an ordinal logistic regression model to examine whether the effects of factors associated with car access vary by rural status, and to what degree. Building upon the literature (Anowar et al. 2014;Basu and Ferreira 2020;Potoglou and Susilo 2008), we selected an ordinal logistic regression model as the appropriate specification for the following reasons: (1) Our dependent variable, car access, is an ordered categorical variable with three levels of responses-carless, car-deficit, and fully equipped; (2) The probabilities by three levels of car access are strictly increasing as the number of categorical responses included increases, known as cumulative probabilities; and (3) The three levels of car access are mutually exclusive, in which the possibility of being in one given response level is dependent on the possibilities of being in other response levels. In an ordinal logistic regression model, the probability of one having a response level J for J = 1, 2, 3, … , j , can be written as: ...
Article
Full-text available
Car access is practically a necessity to get around in rural areas of the United States. Yet approximately 4% of rural residents or 4.3 million people do not have a car. Despite important rural–nonrural differences in the built environment and how people travel, research on rural mobility disparities by car access is limited. This paper addresses this gap by asking and answering the following questions: (1) What are the scope and scale of rural car access in the U.S. and what are the factors associated with not having a car? (2) How do rural carless residents get around relative to their nonrural peers? (3) What are the consequences of not having car access in rural areas? We used a dichotomous rural–nonrural classification to evaluate differences in car access and mobility outcomes between rural and nonrural areas using the U.S. Census Bureau Public Use Microdata Sample and the National Household Travel Survey. The results reveal new evidence on socioeconomic and mobility disparities by car access within and between rural and nonrural areas. Rural carless households earn 64% less and 40% less than their counterparts with full car access and their nonrural counterparts, respectively. Rural Native Americans, Black, and Asian populations have a 2.1 times, 1.3 times, and 1.3 times the odds of being carless, respectively, than their rural white peers, after controlling for other factors. Despite their lack of access to a car, a significantly smaller proportion of rural carless residents bike, walk, or take transit for work trips than their nonrural peers (38.9% vs. 74.6%), which may indicate infeasibility of sustainable modes in some rural areas. Consequently, rural carless residents are 2 times more likely to forsake trips than their nonrural peers due to a lack of transportation options. Our findings highlight rural mobility inequities that carless people face relative to their nonrural peers, which merit attention and tailored policy interventions to improve rural mobility, sustainability, accessibility, and equity.
... The cross-section models, particularly based on discrete choice methods, can be classified into the ordered models (Bhat and Guo, 2007;Maltha et al., 2017) and unordered models (Bhat et al., 2009;Potoglou and Kanaroglou, 2008;Potoglou and Susilo, 2008;Oakil et al., 2014). The ordered logit models reflect the propensity to own a certain ordered number of cars in the family (0 car, 1 car, 2 cars, 3 cars, etc.), while the unordered models typically use the random utility maximization theory and associate certain utility to the choice of owning a car with certain characteristics. ...
... The ordered logit models reflect the propensity to own a certain ordered number of cars in the family (0 car, 1 car, 2 cars, 3 cars, etc.), while the unordered models typically use the random utility maximization theory and associate certain utility to the choice of owning a car with certain characteristics. Some comparisons between the two types of models were conducted and it was concluded that unordered models outperform (Bhat and Pulugurta, 1998;Potoglou and Susilo, 2008). The methods of static Bayesian network (Oakil, 2013) and multiple regressions (Flamm, 2009) were also used to model car-ownership decisions. ...
Article
Individual and household life events are interdependent and influence mobility-related decisions at different levels over time. This paper developed an integrated dynamic model to capture the interdependences among life events, with a special focus on vehicle transactions. Particular attention was paid to the inclusion of vehicles’ characteristics such as the age, fuel type, and size of cars, which are pertinent to emission forecast. A dynamic Bayesian network (DBN), containing individual and household characteristics and latent attitudes toward car ownership and use alongside life events, was employed to study the interdependences. The temporal relationships among life events and lead-lag effects were also captured in the DBN. The longitudinal survey data “the Netherlands Mobility Panel (MPN)” from 2013 to 2018 was used to train and test the DBN. The analysis results confirm the dynamic interdependences between vehicle transactions and other life events and reveal noticeable associations between attitudes and purchase decisions. It is found that several life events (e.g., “Birth of a baby”, “Marital status change”) have concurrent or varied lag-effects on vehicle transaction decisions. The validation indicates that the proposed DBN approach has a high predictive accuracy of vehicle transaction decisions and other life events.
... Notably, the significance of being a high-income household is observed only in the context of the "3 cars" and "4 or more cars per household" alternatives. This suggests that high-income households are more likely to opt for owning more cars, while income may have less impact on the decision to own fewer cars [25] [40]. The residence type variable reveals that single-family detached homes have a stronger influence on auto ownership likelihood compared to single-family attached homes. ...
Conference Paper
Full-text available
Auto ownership behavior is driven by complex relationships influenced by various socioeconomic factors. There also exists overwhelming evidence of racial disparities in auto ownership associated with racial and ethnic residential clustering, termed "automobile mismatch." Yet, information on potential racial and spatial variation in auto ownership preferences is absent from most auto ownership models in practice. This study aims to bridge the research and practical gap in auto ownership models by exploring the additional insights gained when controlling race and space, focusing on Los Angeles County as an ideal case study. Descriptive analyses and auto ownership models for the region were utilized to investigate preferences among racial groups and neighborhoods, revealing systematic variations across racial and spatial dimensions. The composite model, which includes racial and spatial indicators, outperforms the base model, suggesting that the inclusion of race and space explains significantly more information on variations in auto ownership and provides a superior fit to the data. Our findings confirm that auto ownership preferences vary significantly across racial and spatial dimensions, the exclusion of which may skew policy inferences from such models. While future investigations should explore potential underlying reasons for such variations, it is clear that auto ownership models that exclude information on race and space may not be well suited for transportation equity analysis and reflecting potential disparities in auto-based transportation accessibility.
... The basic modeling approach to vehicle ownership is best understood at two levels: the aggregate level, which includes regional, zonal, and national levels, and the household level, which is disaggregated [12]. The disaggregate level is more essential and accurate because it is more effective than aggregate level analysis in gathering data on the causal association between vehicle ownership determinants and automobile ownership levels [13], [14]. ...
... The car-ownership model is a multinomial logit model that estimates for each household whether it owns (i) no household private car, (ii) one car, (iii) two cars, or (iv) three or more cars. Instead of using a multinomial logit model, one could opt for an ordered choice model since the response variable can be considered as ordered However, existing literature shows that multinomial logit models are more appropriate than ordered logit models for studying household car ownership (Bhat and Pulugurta, 1998;Potoglou and Susilo, 2008). For the envisioned application of the model, it must be dependent on the carsharing supply in the living area of the synthetic population. ...
Chapter
As cities grow larger, they often struggle in finding sustainable and liveable mobility solutions to accommodate this growth. Many alternative modes of transport—such as public transport, carsharing systems, bikesharing systems—exist next to private car travel. The effects of expanding those alternatives are often challenging to model. This is in particular the case for small and medium sized cities, which often use straightforward and easy-to-use four-step traffic models. The alternative modes of transport could be modelled using extensive agent-based traffic models. However, these are expensive to make and require a lot of data and expertise. In the context of the EU H2020 project “MOMENTUM”, we developed an intermediate modelling approach that aims to reconcile the user-friendliness of four-step traffic models with the predictive power of agent-based models to investigate the effects of alternative modes of transport. In this paper, we demonstrate the modelling of a policy plan—away from private transport towards durable modes of transport such as shared mobility and public transport—in the city of Leuven, Belgium. We focus particularly on the developed disaggregate car-ownership model, induced demand model, and link-level emission model. It was found that an improved carsharing supply can significantly reduce the car ownership of a city’s households. The largest reduction is seen in households that own several cars and decide they can do with one fewer. These households can use the carsharing system for the occasional trip they would make with the additional car. Moreover, policy measures for the promotion of alternative modes of transport—which might increase the travel times to reach the city for privately owned cars—were found to be able to reduce the city’s mobility-related emissions. In conclusion, we demonstrated that the developed intermediate modelling approach is versatile and applicable to the cities like Leuven, such that they can also account for new modes of transport. The developed models and concepts can help other small- and medium-sized cities to shape their mobility plans.
... Ordered logit model which recognizes the ordinal nature of categories was used. The ordered logit model has been extensively used in transportation literature to model the frequencies of trips (Agyemang-Duah and Hall 1997), satisfaction levels (Bellizzi et al., 2018), number of cars (Potoglou and Susilo 2008), etc. ...
Article
The emergence of eCommerce and online shopping commenced a new episode in human life and changed trading patterns. Online shopping provided access to a broader range of products and facilitated their delivery, which increased demand. To respond to the increased demand, more heavy commercial vehicles need to be on the roads to deliver orders. This is while the road infrastructure is not ready for such a swift shift, and most roads and bridges were planned and constructed during the 19s when online shopping was not coined yet. The continued increase of heavy vehicles on roads can intensify the deterioration of roads and structures such as bridges. Therefore, there is a significant need for an update on new shopping trends, especially changes in people's behavior due to the ongoing Covid-19 pandemic, and to assess if the pandemic permanently changed the trends of in-store and online shopping. This study first examines the Nguyen et al., 2020 data to find the attributes that are significant to online shoppers' behavior. Then a survey is developed to scrutinize Covid-19 effects on the online shopping behavior of users before, during, and after the Covid-19 pandemic. 206 records of data are interpreted through descriptive analysis and discrete choice modeling of users' responses to find the most significant attributes affecting their online shopping behavior. The findings of discrete choice modeling and descriptive analysis support that people tend to go back to stores after the pandemic. The findings of this study show that online and in-store shopping would be balanced after the pandemic and would pursue their normal trends as they were before the pandemic. Based on the findings of this study, it is hard to state that online shopping can vanish in-store shopping due to Covid-19. People still need to go to stores to fulfill their needs for the joy of shopping, interactions with other people, and touching the products they would like to buy. Therefore, transportation stakeholders need to pay special attention to both in-store and online shopping for their planning and operation management of ground transportation infrastructure.
Chapter
Understanding the travel behaviour of road users is important in the development of transportation plans. This chapter present models for household vehicle fleet composition and person-level trip frequencies of both working and non-working adults in Ghana. The study utilizes data from the 2012 Ghana Transport Indicator Database Survey. A Multinomial logit model for vehicle fleet ownership (cars, bicycles, and motorcycles) captured the effects of socio-economic and locational factors. The negative binomial regression models for trip frequencies by mode (walk, bicycle, motorcycle, car, taxi, and bus) highlight important differences in the impacts of socio-economic and location factors on trip frequencies of workers and non-workers. The models collectively demonstrate that factors such age, gender, household composition, income and residential location all have statistically strong impacts on choices about vehicle ownership and trip frequencies by mode. To a large extent, these results appear intuitively reasonable based on common sense expectations and socio-cultural underpinnings of Ghana. The models also highlight the heterogeneity in the vehicle ownership and trip frequency choices across the different parts of the country. Overall, the models from this study can be used to forecast the impacts of changes in socio-economic characteristics of the population on both vehicle ownership and travel demand (trip frequencies) for different parts of Ghana.
Article
Full-text available
This paper offers a critical overview of the full spectrum of household discrete choice–based automobile demand models. Data collection methods, modeling approaches, and the relevant explanatory factors are the primary themes of this review. Furthermore, we examine research methods for assessing the demand for alternative fueled vehicles with an emphasis on stated choices analysis. Overall, this review puts into perspective theoretical assumptions and empirical results for the development of modeling systems capable of assessing market shares and benefits of policy interventions regarding both conventional and alternative fueled vehicle technologies.
Article
It is shown that a strongly consistent estimation procedure for the order of an autoregression can be based on the law of the iterated logarithm for the partial autocorrelations. As compared to other strongly consistent procedures this procedure will underestimate the order to a lesser degree.
Article
Vehicle ownership plays a key role in determining the number of trips made by a household and the modes that the household will use to make those trips. Predicting vehicle ownership is a challenge due to the complexity of relationships affecting the choice of vehicles for a particular household. The income of the household, the accessibility of activities by alternative modes, and the family makeup of a household all affect the choice of the number of vehicles. The Honolulu metropolitan statistical area, which encompasses the island of Oahu, is the primary study area. On Oahu, vehicle ownership varies widely, from the dense neighborhoods in Waikiki to the semirural areas on the fringe of the region. Consequently, the vehicle-ownership model will be a crucial component of the new travel forecasting models for the island. Using 1995 Oahu household interview survey data, a vehicle-ownership model that takes into account income, accessibility, and household family structure has been estimated and found to predict rather accurately the shares in each vehicle-ownership class by geography, household size, and number of workers. In addition, 1990 census data have been used to test a variation of the vehicle-ownership model on several other American cities such as Atlanta, Kansas City, San Francisco, and New York. The resulting models for these regions parallel the Honolulu vehicle-ownership model very closely in structure and statistical significance, thereby demonstrating the model structures' transferability between regions.
Article
A mathematical model developed to predict automobile ownership for individual households residing in New York City is presented. This effort is distinguished from previous disaggregate household-level automobile ownership models primarily by the use of ordered probit models rather than the commonly used multinomial logit (MNL) and sequential logit (SL) models. When the dependent variable involves ordinal categorical data (in this case, automobile ownership level--zero automobiles, one automobile, two automobiles, and three or more automobiles), the ordered probit model will discern unequal differences between ordinal categories in the dependent variable, the MNL model will treat categories as independent choice alternatives, and the SL model (a product of binary logits) will assume independence of the error terms across all binary choices. The modeling approach was based on a behavioral analysis that explained the factors influencing household automobile ownership decisions in a highly urbanized environment. In addition to socioeconomic variables, transportation and land use-related measures were developed and used to test the sensitivity of household automobile ownership choice to transit accessibility, traffic congestion, parking cost and availability, and levels of access to opportunity sites through nonmotorized transportation. The estimation results uncover important interactions between socioeconomic- and location-related elements and automobile ownership. Findings provide exploratory methodological and empirical evidence that could lead to an approach to predicting the change in household automobile ownership as a result of changes in future socioeconomic conditions and transportation and land use scenarios.
Article
These Guidelines describe how transit agencies, metropolitan planning organizations, and state departments of transportation can act today to initiate or expand their analytical tools for integrated land use-transportation planning. The Guidelines are intended for the general reader having an interest in the effects of transit on land use. The Guidelines describe currently available integrated models, the characteristics of an "ideal" integrated model, and steps that a planning organization should take in order to support and expand such modeling capability. A more detailed Final Report is available for the practitioner on the TCRP website as Web Document 9 (www4.nas.edu/trb/crp.nsf).