You are on page 1of 11

Comparative analysis of trip generation models:

results using home-based work trips in the


Seoul metropolitan area
Justin S. Chang*, Dongjae Jung, Jaekyung Kim and Taeseok Kang
This paper compares the performance of trip generation models. Trip generation estimates the
number of trips to and from a traffic analysis zone. This process is the first stage of the conventional
four-step travel forecasting framework. Although many approaches have been suggested for
this step, regression and category analyses have been widely applied. The two methods have
generated an acceptable level of performance from the perspective of transport planning. Critical
problems, however, have also been observed. In the regression analysis, trip rates are treated as
continuous variables that can be negative, which is obviously unrealistic. Furthermore, the method
does not incorporate traveler behavior. For the category analysis, its arbitrary way of choosing
independent variables and their strata has drawn critiques. The cell-by-cell calculation in this
method also increases the concerns about unreliable estimation of trip rates. Censored regression,
count data, and discrete choice models have been visited for the alternative of regression approach
while the multiple classification method has been conceived for the substitute of the category
analysis. A systematic examination of the performance among the models has not been discussed
sufficiently yet, which is the motive of this paper. Six representative models regression, tobit,
Poisson, ordered logit, category, and multiple classification analyses were applied to the homebased work trips in the Seoul metropolitan area. Cross-validation and back-casting were the key for
checking the performance among the models. In this process, the measures of correlation,
variance, and coincidence were compared. The category-type model was superior in overall
performance.
Keywords: Trip generation, Regression model, Tobit model, Poisson model, Ordered logit model, Cross-classification, Multiple classification analysis, Validation

*Corresponding author, email jsc@snu.ac.kr

The trip generation estimation procedure employs


mathematical models that associate each purpose with
demographic characteristics of the TAZ, such as population, households, employment, vehicle ownership, and
income. Current information on these variables may be
obtained from household surveys or census reports. Future
information, on the other hand, is derived from projections.
Early versions of trip generation were mainly based on
zonal aggregation approaches. However, the aggregate
models lack the context of travelers behavior. Householdbased schemes, thus, are more common in current practice,
even though they require an additional process of obtaining
zone-level totals (Papacostas and Prevedouros, 2001; 351
353).
Linear regression (hereafter regression) and category
analyses are the representative methodologies used for this
step. They have widely been applied to empirical studies

2014 W. S. Maney & Son Ltd


MORE OpenChoice articles are open access and distributed under the terms of
the Creative Commons Attribution Non-Commercial License 3.0
Received 15 October 2013; accepted 17 January 2014
DOI 10.1179/1942787514Y.0000000011

Transportation Letters: the International


Journal of Transportation Research

Introduction
This paper compares the performance of trip generation
models. Trip generation, which is the first phase of
conventional four-step travel forecasting framework, estimates the number of trips to and from each traffic analysis
zone (TAZ) for various purposes. This trip-based traditional approach is still the standard practice for most
strategic transport planning, even though advanced
approaches like tour- and activity-based models explore
more realistic representations of behavior in travel demand
studies (McNally and Rindt, 2007; Donnelly et al., 2010).

Department of Environmental Planning, Seoul National University,


1 Gwanak-ro, Gwanak-gu, Seoul 151-742, Korea

78

2014

VOL

NO

Chang et al.

and may have shown acceptable performance from the


planning perspective.
However, there are also limitations of these traditional
frameworks. For regression-based trip generation models,
three typical drawbacks have been observed. First, the
number of trips is treated as a continuous random variable
though it is a discrete one (Barmby and Doornik, 1989;
Ma and Goulias, 1999; Wallace et al., 1999; Jang, 2005;
Schmocker et al., 2005; Badoe, 2007; Roorda et al., 2010;
Lim and Srinivasan, 2011). People, for example, can make
two trips per day; people cannot make 1?7 trips per day.
Second, the dependent variable may take on negative
values due to the assumption of normal distribution for
the disturbance of trip rates (Barmby and Doornik, 1989;
Cotrus et al., 2005; Ma and Goulias, 1999; Wallace et al.,
1999; Jang, 2005; Badoe, 2007). In trip generation, the
dependent variable is zero for a significant fraction of the
observations. For those who make trips, the travel
demand can be measured; but for those who do not, the
spatial interaction cannot be recorded and is set equal to
zero. Namely, although the data for trip generation are
censored, the regression model cannot address the nature.
Finally, the model does not represent traveler behavior
theory because it simply matches a statistical relationship
between the dependent variable and a set of independent
variables (Schmocker et al., 2005; Badoe, 2007; Roorda
et al., 2010; Lim and Srinivasan, 2011). It is difficult to
observe travelers behavioral mechanisms in trip-making,
such as utility maximization and cost minimization. The
category analysis may have useful advantages over
regression-type trip generation models (Stopher and
McDonald, 1983). The technique is independent upon
the zonal system of the study area; no assumptions are
required about the shape of the relationship between the
trip rate and explanatory variables; it represents classspecific behavior; and it does not permit extrapolation
beyond its calibration strata, although the lowest or
highest class of a variable can be open-ended. However,
the model also suffers from two broad limitations (Stopher
and McDonald, 1983). First, a nave approach in choosing
independent variables and their strata for classification is
not statistically justifiable but only empirically acceptable.
The method bears no statistical goodness-of-fit measures,
so the calibration cannot be verified. Second, the cell-bycell calculation reduces the reliability of cell values. In
particular, the uncertainty increases when there are cells
with small samples and/or large variances. This calculation
mechanism also requires large sample sizes, which incurs
much cost and time (Ortuzar and Willumsen, 2011;
158).
There has been research to overcome and/or mitigate
these limitations. Censored regression models such as tobit
analysis have been used to block the potential negative
values on trip generation rates (Cotrus et al., 2005).
Poisson and negative binomial models have been used to
try to address the integer nature of the dependent variable
(Barmby and Doornik, 1989; Ma and Goulias, 1999;
Wallace et al., 1999; Jang, 2005; Badoe, 2007). These
count data models can also confine the figure of the trip
rate greater than or equal to zero. Ordered logit and probit
models can be understood as generalized frameworks in

Comparative analysis of trip generation models

regression-based trip generation modeling since the


discrete choice methods are, in principle, free from the
three limitations of the likelihood of negative trip rates,
the continuous dependent variable, and the lack of
incorporation of traveler behavior theory (Sheffi, 1979;
Agyemang-Duah and Hall, 1997; Schmocker et al., 2005;
Badoe, 2007; Roorda et al., 2010). Multiple classification
analysis is also found in the literature as an alternative
approach to a simple category model for person trips
(Stopher and McDonald, 1983) and for freight movements
(Bastida and Holguin-Veras, 2009). This model adopts a
statistically justifiable approach based on analysis of
variance (ANOVA). The cell value is estimated with the
grand and group means (Stopher and McDonald, 1983).
As reviewed, there have been studies to alleviate the
limitations of conventional trip generation models. Also,
some literature has compared two models (e.g., Cotrus
et al., 2005) or more (e.g. Badoe, 2007; Lim and Srinivasan,
2011). The comparative studies, however, have mostly
focused on the estimation results and a rough validation,
normally with current datasets. On the other hand, a
systematic validation is explored in this study to compare
the performance of models, incorporating both crossvalidation and the historical method in the form of backcasting. It should be noted that the model comparison does
not directly ascertain whether the forecast response is
correct, but does assess whether it is reasonable or
explainable given what is known about traveler behavior.
The next section describes the scope and design of this
study. Subsequent two sections deal with the theoretical
basis of the trip generation models which are tested in this
paper. The process and results of the empirical study then
follow. The section includes data used, estimation results,
and performance comparison of the models. Finally,
concluding remarks are given.

Scope and design


Figure 1 summarizes the critiques on the conventional trip
generation models as discussed in the introductory section.
The diagram also shows the alternative approaches and
their representative models, for which performance is
tested in this paper.
In fact there are other models for each group of
alternatives. The ordered probit (Schmocker et al., 2005;
Roorda et al., 2010; Lim and Srinivasan, 2011), the
negative binomial (Wallace et al., 1999; Badoe, 2007; Lim
and Srinivasan, 2011), and the truncated normal (Badoe,
2007) models have been applied from the groups of
discrete choice, count data, and censored regression
frameworks, respectively. However, they are less common
methodologies than those chosen in Figure 1.
It should also be noted that there are approaches which
are not included in this comparative study. Highly
sophisticated techniques such as the count data model
with zero inflation (Jang, 2005) and the two-limit tobit
version are not included in this study. The binary stop-go
model (Daly, 1997) is not considered either. There may be
others which are not enumerated in this paper. Indeed,
these advanced models are associated with appealing

Transportation Letters: the International Journal of Transportation Research

2014

VOL

NO

79

Chang et al.

Comparative analysis of trip generation models

1 Conventional approaches and their alternatives

theoretical structures. However, their corresponding


empirical performances have not generally been verified.

qn ~bT xn zen , n~1,2,    ,N


qn ~0

Regression-type models

qn ~qn

if qn 0
if

(2)

qn w0

Regression model
Regression analysis in trip generation functionalizes the
relationship between trip generation rates, or the dependent variable, and a set of independent variables
qn ~bT xn zen ,

n~1,2,    ,N

(1)

where q is the trip generation rate, n indexes the nth


observation (which is households in this study), b is the
vector of parameters that should be estimated, x is the
vector of independent variables, and e is a random
disturbance.
Regression models require rather strong assumptions.
Four representative issues are that the linear functional
form is demanded; the model parameters should be
identifiable, namely the full rank condition should be
satisfied; the data on the independent variables are nonstochastic; and the disturbances are normally distributed,
with zero mean and constant (or homoscedastic) and
uncorrelated (or independent) variance (Greene, 2000;
213223).

The result of the partial derivative with respect to any


independent variable cannot be the marginal effect of the
variable, since q is unobserved. The marginal effect in this
kind of model, or the case with censoring at zero and
normally distributed disturbances, is given by
 T 
b xn
(3)
bW
s
where W(?) is a normal cumulative distribution function
and s is the standard deviation of the error term (Greene,
2000; 905926).

Poisson model
Trip rates, or the dependent variable, clearly show a
discrete nature. In addition, a very large proportion of trip
rates by households are zeros or small values. Regression
analyses based on least squares hold limitations for
dealing with these characteristics. A count data model
such as a Poisson regression can be an alternative
framework

Tobit model
Only the part of the distribution above q 50 is relevant to
the modeling for trip generation. Conventional regression
methods fail to account for the qualitative difference
between limit (zero) and non-limit (continuous) observations. The tobit model is a useful tool to deal with these
censored data. When data are censored, the distribution
that applies to the sample data is a mixture of discrete and
continuous distributions (Cotrus et al., 2005). To analyze
this distribution, a new random variable q which is
transformed from the original q is defined

80


 e{ln lqnn
, n~1,2,    ,N
P qn ~qn ~
qn !

(4)

where P(?) is the probability, the log-linear function is


normally applied to ln, namely ln ln5bTxn, and the
partial effects in this non-linear regression model are given
by lnb (Greene, 2000; 880893).
The Poisson model, however, shows a critical limitation:
it assumes that the conditional mean and variance
functions are equal. Overdispersion is particularly problematic (Barmby and Doornik, 1989; Ma and Goulias, 1999;

Transportation Letters: the International Journal of Transportation Research

2014

VOL

NO

Chang et al.

Wallace et al., 1999; Jang, 2005; Badoe, 2007). If


overdispersion is detected, the negative binomial model,
which considers the variance to differ from the mean, is
usually considered to apply. However, the data used for
this study are not very concerned with this problem, as
shown in Table 2.

Ordered logit model


The observed number of trips made by households is
ordered. The count data model explains these rates using a
definite point value. However, it would be more natural to
think that a decision-maker has some level of utility associated with trip-making. When the level of utility is above
some cutoff values, it is understood to lead to trip-making.
The ordered logit model specifies trip rates as a set of
ordered alternatives. It is not difficult to regard that one
alternative is similar to those close to it and less similar to
those further away (Train, 2009; 159164)
qn ~0 if qn h0
qn ~1 if h0 vqn h1
..
.
qn ~k

(5)

if hk1 vqn hk
..
.

where hs are cutoff points for the possible responses and


qn has the similar specification to the other models in this
study, namely qn ~bT xn zen .
Provided that en is distributed logistic, the probability of
household n making k trips is given by




Pqn ~k~F hk {bT xn {F hk{1 {bT xn
(6)
where F(?) is the cumulative distribution function of en.
The marginal effects of the ordered logit model is given by
 



F hk{1 {bT xn {F hk {bT xn b
(7)
where F9(?) is the logistic density function and thus the
effects represent the rate of change in the probability of tripmakings associated with a unit change in the independent
variables, holding all other variables constant.

Cross-classification-type models
Category analysis
The category analysis, or the cross-classification method,
would be the most extensively used approach for trip
generation. In this framework, household types are
classified according to a set of categories that are highly
correlated with trip-making. The dependent variable is
assumed to be continuous and two or three explanatory
variables, each broken into three or four discrete levels,
are usually applied. The independent variables can include
household size, car ownership, household income, and
some measures of land-use.
Mathematically, each household category constitutes a
cell in the cross tabulation. The average trip rate in a cell is
then calculated by simple algebra

Comparative analysis of trip generation models

Qh
q- h ~
H

(8)

where q- h is the average number of trips made by


household type h (e.g., the one car and two-worker
household); Qh and Hh are the total number of trips and
households observed for type h, respectively.

Multiple classification
The conventional category analysis suffers from two
important methodological drawbacks: a non-statistical basis
and cell-by-cell calculation. The multiple classification
method can overcome these disadvantages. A statistically
justifiable approach based on a series of ANOVAs is applied
to select independent variables and their strata. First, oneway ANOVAs between trip rates and each candidate
variable are used to find the best grouping for each variable.
In this process, measures such as an F statistic and R2 can be
used. Once statistically significant independent variables
have been identified, multi-way ANOVAs between trip
rates and two or three candidate variables are applied. It is
an extensive trial-and-error procedure to examine the best
classification scheme. The eta-square g2, which is the ratio
of the sum of squares of the candidate variable to the
corrected total in the ANOVA output, can be used to
determine more contributable variables (Stopher and
McDonald, 1983). A grand-group mean approach is used
to estimate the cell value. The grand mean is calculated over
the entire sample, while the group mean is computed from
the row and column sums of the category table. A cell value
is found by adding the deviations of the cell to the grand
mean; this process is different from the cell-by-cell
calculation in the conventional category analysis (Stopher
and McDonald, 1983).
There are two drawbacks to multiple classification analysis. The ANOVA-based statistically justifiable approach
requires extensive trial-and-error procedures with no claim
of optimality. Its empirical acceptance is subject to low
efficiency in selecting good classification structures. The
grand-group mean calculation also loses the class-to-class
relationship between independent variables since one
variable is computed over all classes of the other variable
(Stopher and McDonald, 1983).

Method for model comparison


Schemes
The goodness-of-fit for each model are examined based on
several procedures: the signs of the variables are checked
against engineering judgment and corresponding expectations; a t-test based on the standard error of the estimate
is applied to inspect the statistical significance of each
variable; a statistic which is used for exploring the
significance of the entire model is estimated, such as the
F-statistic for the regression analysis and the likelihood
ratio test for the ordered logit model; and informal
indicators such as the coefficient of determination R2 and
McFaddens r2 are examined.
However, the comparison between models is not
straightforward. The magnitude of the estimated coefficients cannot be meaningfully compared between the

Transportation Letters: the International Journal of Transportation Research

2014

VOL

NO

81

Chang et al.

Comparative analysis of trip generation models

models since the scales are different. The same holds true
for the standard goodness-of-fit measures such as the Ftest, R2, and t-test.
The performance comparison between models is conducted with validation. Validation tests a models ability to
predict future behavior. It requires comparing the model
output with information other than that used in estimating
or calibrating the model. Namely, the model output is
compared with observed travel data. Two ways of checking
model performance are considered in this paper. They are
the historical method and cross-validation.
Cross-validation is a statistical method for validating a
predictive model. Subsets of the data are held out for use
as validating sets; a model is fit to the remaining data and
used to predict for the validation set. Averaging the
quality of the predictions across the validation sets yields
an overall measure of prediction accuracy.
There are three distinct forms in the cross-validation.
Table 1 summarizes the advantages and limitations of the
cross-validation techniques. The simplest type would be
the hold-out technique, which is adopted in this study. It
divides observations into two subsets: one is for estimation and calibration, and the other is for validation.
Another form of cross-validation leaves out a single
observation at a time; this is similar to the jackknife
technique in the resampling. Lastly, the K-fold crossvalidation technique splits the data into K subsets; each is
held out in turn as the validation set (Picard and Cook,
1984; Kohavi, 1995).
The hold-out method is easy to apply, but there are few
reliable rules for classifying samples into the estimation
and validation sets. If data for estimation and validation
are sufficient enough as in this study, this issue does not
empirically matter much. The leave-one-out technique
may incur minimal errors because it simulates total
sample-size times, leaving just one observation out in
every trial. However, this method is inapplicable to large
sized datasets. The K-fold method requires acceptable
computation time and may generate reasonable errors in
the validation. However, this scheme also suffers from the
arbitrariness in determining the number K.
The historical technique is either forecasting, in which a
prior-year model is used to forecast current travel that is
then compared with actual current travel, or back-casting,
in which a current year model is used to estimate travel for a
prior year that is then compared with actual travel in the
prior year. The literature establishes little consensus on the
better of the two: the choice is fundamentally datadependent (Committee for Determination of the State of
the Practice in Metropolitan Area Travel Forecasting,
2007). Using both trials, of course, would be good practice.
In this paper, back-casting is used due to data availability.

Measures
It would be desirable to have some quantitative measures
for validation. In this paper, measures of correlation,
variance, and coincidence are considered (Cambridge
Systematics, Inc., 2010).
Correlation refers to a statistical relationship between
the observed and estimated trip rates. The closeness can be
expressed numerically by the correlation coefficient r
sq q^
(9)
r~ n n
sqn sq^n
where ^
qn refers to the estimated value of qn , sqn q^n is the
qn , and sqn and sq^n are the
covariance between qn and ^

qn , respectively.
standard deviations of qn and ^
Root-mean-square error (RMSE) is a measure of the
differences between values predicted by a model and those
actually observed
v
u N
u1 X
2
^
(10)
qn {qn
RMSE~t
N n~1
Though RMSE is a good index of model accuracy, it is
vulnerable to the scale problem. Per cent RMSE is an
alternative because it is the normalized version of RMSE,
removing scaling effects
v
u N 

u1 X ^
qn {qn 2
(11)
% RMSE~t
N n~1
qn
Low RMSEs and high values of the correlation
coefficient are only one desirable measure of fit. Another
important criterion is how the model simulates the pattern
of the data. Namely, the distribution of estimated trip
frequency should be analogous to the shape in the data.
Thus, the ability to duplicate turning points or rapid
changes in the data is an important criterion for model
evaluation. The coincidence plot and its corresponding
ratio are a useful tool to address this issue. The plot
compares estimated and observed trip frequency distributions. The ratio measures the per cent of area that
coincides. This rate lies between zero and one, where zero
indicates two disjoint distributions and one means an
identical pattern (Cambridge Systematics, Inc., 2010)
(
!)
^k
P
H
Hk
min P
,P
^
k
k Hk
k Hk
(
!)
(12)
^k
P
H
Hk
max P
,P
^
k
k Hk
k Hk

Table 1 Advantages and limitations of the cross-validation methods

Hold-out method
K-fold method
Leave-one-out method

82

Advantages

Limitations

Easy to apply
Known to involve reasonable levels of errors
Acceptable computational needs
Minimal errors incurred in principle

Little reliable rule for the sample classification


Little reliable rule for determining the number K

Transportation Letters: the International Journal of Transportation Research

Too much computational burden

2014

VOL

NO

Chang et al.

^ k are the total number of households


where Hk and H
observed and estimated at trip count k, respectively.

Empirical analysis
Data
The study area is the Seoul metropolitan region in Korea.
The geographical sector includes Gyeonggi province;
Incheon, which is the fourth largest city in Korea; and
Seoul, the largest city and capital of the country. Subregions establish the functional relationships in terms of
lifestyles, economic activities, urbanization, and land-uses.
There are two broad sets of data used in this study. The
data for household characteristics and trip generation rates
are based on the household travel diary survey of the Seoul
metropolitan area. The Korea Regional Development
Total Information System (REDIS) database supplies the
regional characteristics that affect home-based work trips.
The Korea Regional Development Total Information
System is the official statistical database of the Korea
Presidential Committee on Regional Development. It
collects more than 300 regional statistics and has provided
the data to the public since 2008. This study has used four
types of data: demographics, regional economy, transport
system, and land-use.
The travel diary survey was started in 2002 by the code
prescribed in the Korea Intermodal Surface Transportation
Efficiency Act. This study extracted home-based work trips
from data surveyed in 2002 and 2006. The data of 211 564
households in 2006 and those of 159 068 households in
2002 were collected, as shown in Table 2. Two-thirds of the
data from 2006 were used for the estimation while the
remaining data from 2006 and all the samples in 2002 were
applied for the validation. Since the sample size of the 5 or
more trip rate is too minimal compared to those of the
others, the category was excluded in this study. Thus the
sample size in Table 5 is 148 018 not 148 091.

Estimation results
Table 3 shows the independent variables considered in the
study. Three kinds of household characteristics and four
sets of regional characteristics were considered.
The specifications for regression, tobit, Poisson, and
ordered logit models are shown in Table 4. These are the

Comparative analysis of trip generation models

result of trial-and-error experiments for the modeling of


trip generation, finding an empirically best fit for each
approach. The process is subject to arbitrariness, but it
cannot be avoided in comparing different models; note
that the models tested have different mathematical
structures. However, it should be stressed that the four
models have the same independent variables through the
heuristic method for specification, which is helpful for a
fair comparison across the models tested.
Table 5 shows the estimation results for regression,
tobit, Poisson, and ordered logit models. The results show
satisfactory goodness-of-fit. The formal significance
indices of F- and the likelihood ratio statistics indicate
that the null hypothesis that all the parameters are zero
can be rejected at the 0?01 level of significance. The values
of R2 and r2, though they are an informal goodness-of-fit
index, are also at an empirically acceptable level.
However, the r2 of the Poisson model is relatively low.
Not just the overall goodness-to-fit but also the
reasonableness of each parameter should be checked.
Since the marginal effects of independent variables are not
equal to the coefficients in the tobit, Poisson, and ordered
logit models, the effects are also given. The figures in the
parentheses of Table 5 are the partial effects of parameters
of the tobit and Poisson models while those for the
ordered logit model are shown in Table 6. Even though
the discussion regarding the partial effects is indirect to the
purpose of this paper, the effects are useful for examining
sensitivity to the explanatory variables.
All the coefficients in the model estimates have the
expected signs. The variable number of employees is the
direct explanatory factor for the home-based work trips. It
is common sense that more commuters make more
commuting trips. This element, however, is not the only
cause for trip generation. The variables household
income and car ownership are the independent variables
from the group of household characteristics. They reflect
the degree of economic activities by household. Namely, it
can be expected that a trip-maker with vehicles and higher
income would actively participate in economic activities.
This results in more spatial interactions in the form of
commuting trips. The variable number of subway lines is
the factor of regional characteristics in terms of the
transport system. Since the study area is the urbanized

Table 2 Number of households by trip rate


Data in 2006

Data in 2002

Trip rate

Total

Estimation set

Validation set

Validation set

0
1
2
3
4
5 or more
Total
Average
Standard deviation

43 521 (20.57)
115 358 (54.53)
43 714 (20.66)
7630 (3.61)
1234 (0.58)
107 (0.05)
211 564 (100.00)
1.1130
0.7702

30 455 (20.57)
80 738 (54.52)
30 616 (20.67)
5337 (3.60)
872 (0.59)
73 (0.05)
148 091 (100.00)
1.0928
0.7804

13 066 (20.59)
34 620 (54.54)
13 098 (20.64)
2293 (3.61)
362 (0.57)
34 (0.05)
63 473 (100.00)
1.0920
0.7799

29 237 (18.38)
91 139 (57.30)
31 504 (19.81)
5961 (3.75)
1126 (0.71)
101 (0.06)
159 068 (100.00)
1.0926
0.7802

The number in parentheses refers to the percentage; return-home trips are excluded since origin-destination-based data are only available
in 2002 while production-attraction-based data are also built in 2006.

Transportation Letters: the International Journal of Transportation Research

2014

VOL

NO

83

Chang et al.

Comparative analysis of trip generation models

Table 3 Independent variables considered

Household
characteristics

Variable type

Variables considered

Household composition

Number of telecommuters
Number of employees
Number of household members
Number of preschool children
Household income
House ownership
Car ownership
Number of cars owned
Subway availability
Number of household members with driving licenses
Total population
Population density
Population moved in
Population moved out
Annual expenditures
Annual expenditures per capita
Annual expenditures per capita per area
Annual tax revenues
Annual local tax revenues
Annual local tax per capita
Fiscal self-reliance ratio
Number of car registered
Number of car registered per capita
Number of subway stations
Number of subway lines
Number of bus routes
Road density
Gross area
Residential area
Commercial area
Industrial area
Green area
Number of employees
Number of employees per area
Number of companies
Number of companies per area

Income level
Transportation

Regional
characteristics

Demographics

Regional economy

Transport system

Land-use

Seoul metropolitan region, this geographic sector suffers


from the chronic congestion that other large cities normally
do. Hence, households would prefer their residential
locations close to public transportation. In particular, rail
transport is known to be the most reliable travel mode in
terms of travel time. Thus, the mandatory journey has a
positive causality with transit proximity. The length of
subway lines can also be considered as an alternative for the

proximity variable. However, the aggregate measure would


be restrictive to explain the number of options in rail travels
that trip-makers face in their daily lives. The variables
population density, annual expenditures per capita per
area, and number of companies per area from the categories of demographics, regional economy, and land-use
are directly or indirectly related to employment opportunity. More companies can mean more jobs, which require

Table 4 The specication for regression, tobit, Poisson, and ordered logit models
Variable
Dependent variable
Trip rates
Independent variables
Number of employees
Household income

Car ownership
Population density
Annual expenditures per capita per area
Number of subway lines
Number of companies per area

84

Specification

Number of trips per household


Number of employees per household
1 if household income is below 1000 K won
2 if household income ranges 10002000 K won
3 if household income ranges 20003000 K won
4 if household income is over 3000 K won
1 if a household has cars; 0 otherwise
1000 persons per square kilometer
1000 K won per person per square kilometer
Number of lines within walking distance
1000 companies per square kilometer

Transportation Letters: the International Journal of Transportation Research

2014

VOL

NO

Chang et al.

Comparative analysis of trip generation models

Table 5 Estimation results for regression, tobit, Poisson, and ordered logit models

Constant
Number of employees
Household income
Car ownership
Population density
Annual expenditures per capita per area
Number of subway lines
Number of companies per area

Regression

Tobit

Poisson

0.0098
0.5394***
0.0947***
0.0072*
0.0022***
0.3014***
0.0091***
0.0551***

20.3352 (20.3073)***
0.6166 (0.5652)***
0.1303 (0.1194)***
0.0323 (0.0296)***
0.0031 (0.0028)***
0.2972 (0.2724)***
0.0101 (0.0093)***
0.074 (0.0678)***
0.7248 (0.6644)***
0.9167

20.9518
0.4208
0.1004
0.0276
0.0030
0.2878
0.0068
0.0432

148 018
MLE

148 018
MLE

3.5167***
6.4983***
9.0627***
148 018
MLE

75 740.6***

29 981.8***

77 142.6***

Scale factor for marginal effects

Ordered logit
(21.0383)***
(0.4591)***
(0.1095)***
(0.0301)***
(0.0033)***
(0.3139)**
(0.0075)
(0.0471)***

21.8605***
1.9776***
0.3307***
0.0469***
0.0063***
0.9318***
0.0314***
0.1880***

1.0909

h1
h2
h3

Sample size
Estimator
F-statistic

^

{2 L0{L b

148 018
OLS
68 367.6***

R2

0.4005

0.0828
0.2302

^

* 10% significance level; ** 5% significance level; *** 1% significance level; {2 L0{L b is the likelihood ratio statistic where L(0) is
^

the value of the log likelihood function when all parameters are zero and L b is that at its maximum.
2
r

0.1765

more people. As a result, an area becomes denser. A local


government normally earns more taxes in a region with
large population, resulting in governments spending larger
budgets.
The term constant, on the other hand, should be
understood more carefully. It represents the disturbance in
the model formulation. Thus, it should be read as the
collective effects of the independent variables that are not
included in the specification. Hence, the positive sign, as in
the regression model, would match to the statistical
background.
However, not all the coefficient estimates are significantly different from zero at the usual 5 or 1% levels of
significance. Note that for most coefficients, the null
hypothesis that the true value is zero can be rejected at the
1% significance level, but some others are not statistically
different from zero at the same significance level.
However, the inability to reject the hypothesis that some
coefficients are zero at a particular significance level does
not imply that the hypothesis must be accepted.
Table 7 shows the results of the traditional category
analysis. The selection and stratification of independent
variables of this scheme fundamentally follows an ad hoc
procedure. The choice of explanatory factors comes from
the result of the multiple classification analysis. The levels

of household income are the same as the specification of


the previous four models. Also, a simple increment for the
grouping has been applied as in a usual category scheme.
These designs allow for consistent performance comparison between the models tested.
Table 8 is the summary of the one-way ANOVAs for the
grouping of the independent variables for multi-category
Table 7 Average daily commuting trips per household by
cross-classication
Number of employees
Population Household
density
income
Low

Medium

High

01000 K won
10002000 K won
20003000 K won
3000 K won or more
01000 K won
10002000 K won
20003000 K won
3000 K won or more
01000 K won
10002000 K won
20003000 K won
3000 K won or more

3 or more

0.26
0.69
0.87
0.90
0.08
0.26
0.32
0.41
0.11
0.33
0.39
0.47

0.47
0.76
0.84
0.86
0.73
0.85
0.89
0.90
0.74
0.85
0.91
0.92

0.76
1.27
1.48
1.58
1.35
1.54
1.61
1.66
1.36
1.53
1.60
1.69

1.37
1.52
1.98
2.27
1.83
2.11
2.22
2.39
1.93
2.18
2.33
2.51

Table 6 Marginal effects of ordered logit model


Independent variable

q50

q51

q52

q53

q54

Number of employees
Household income
Car ownership
Population density
Annual expenditures per capita per area
Number of subway lines
Number of companies per area

20.20249
20.03386
20.00480
20.00065
20.09540
20.00322
20.01925

20.09550
20.01597
20.00227
20.00031
20.04500
20.00152
20.00908

0.27576
0.04611
0.00654
0.00088
0.12993
0.00438
0.02621

0.02048
0.00342
0.00049
0.00007
0.00965
0.00033
0.00195

0.00175
0.00029
0.00004
0.00001
0.00082
0.00003
0.00017

Transportation Letters: the International Journal of Transportation Research

2014

VOL

NO

85

Chang et al.

Comparative analysis of trip generation models

Table 8 The result of one-way ANOVAs for the multiple classication analysis
Candidate variable

Stratification

F-statistic

Number of employees
Household income
Car ownership
Population density
Annual expenditures per capita per area
Number of subway lines
Number of companies per area

0, 1, 2, and 3 or more
01000 K won, 10003000 K won, and 3000 K won or more
0 and 1 or more
Low (05000), medium (500018 000), and high (18 000 or more)
0100, 100200, 200300, and 300 or more
0, 1, 2, 3, and 4
0500, 5001000, and 1000 or more

31541.922***
4647.556***
1306.429***
722.164***
58249.731***
33.762***
285.863***

Population density, annual expenditures per capita per area, and number of companies per area are represented by persons per square
kilometer, 1000 won per capita, and companies per square kilometer, respectively; and *** 1% significance level.

analysis. All the explanatory variables considered in the


regression, tobit, Poisson, and ordered logit models also
show statistical significance in this multivariate analysis.
In general, two or three independent variables can be
considered for this analysis in the form of two- or threeway ANOVAs, respectively. Since seven candidate variables are all statistically significant, the scheme requires a
series of ANOVAs with a trial-and-error basis. The details
of the extensive ANOVA runs are not provided in this
paper: the result of the empirically best run, based on a
three-way ANOVA, is given in Table 9.
The cell values in the conventional and multiple crossclassifications generally match prior expectations. Namely,
the trip rate increases as the income is higher and the
number of employees grows. Two interesting observations, however, should be noted. First, two couples of cells
that are shaded in the traditional category analysis
contradict common sense. Intuitively, everything else
being equal, the one-employee household can be expected
to make more trips than the non-employee household.
However, the outcome is the reverse. Meanwhile, the
reason in which non-employee households make trips is
that there can be employers in the household who
contribute to the number of commuting trips. Second,
the cell values in the conventional model are similar
between the medium and high density areas. This may
suggest that population density can be classified into low
and a combined medium and high. It may also imply that
the other independent variables are not very well stratified.
Unfortunately, there is little systematic way to identify the
Table 9 Average daily commuting trips per household by
multiple classication
Number of employees
Population Household
density
income
Low

Medium

High

F-statistic
2
R

86

01000 K won
10003000 K won
3000 K won or more
01000 K won
10003000 K won
3000 K won or more
01000 K won
10003000 K won
3000 K won or more
3237.538***
0.434

3 or more

0.07
0.52
0.66
0.18
0.36
0.43
0.31
0.50
0.59

0.10
0.54
0.68
0.73
0.92
0.99
0.82
1.00
1.10

0.69
1.13
1.27
1.45
1.64
1.71
1.53
1.71
1.81

1.18
1.63
1.77
2.09
2.27
2.34
2.25
2.44
2.53

cause in the conventional category analysis. The multiple


classification has generated the merged group of the factor
household income through the ANOVA procedure; note
that there are four and three strata in the category and
multiple classification analyses, respectively.

Performance comparison
The performance comparison between the models is
conducted based on validation. As stated in the methodological section, both cross-validation and back-casting
were adopted in this paper. The measures of correlation,
variance, and coincidence were utilized for the quantitative assessment of this process.
Since validation is basically represented by the comparison between the observed and modeled values, the first
task is to determine estimated trip rates for each model.
This is straightforward in the case of discrete choice
models. The continuous and count data models, however,
require an auxiliary step to convert model outputs to
integer approximations. The category-type models also
need this additional phase. The procedure adopted in this
paper is to round off the values to the nearest integer. This
process, however, inevitably involves biases.
Table 10 summarizes the results of validation. In the
cross-validation, the models show similar RMSEs while the
category-type approaches achieve slightly better performance. This trend is also observed from the % RMSE
measure. This result is not surprising because the two
indices represent fundamentally the same performance, or
variance, even though % RMSE eliminates the scale
problem of RMSE. The differences in model performance
are noticeable when viewing the correlation indicator.
Again, the category-type methods accomplish a better
goodness-of-fit while the Poisson model records the worst.
A comparable finding is seen in the coincidence index.
Back-casting results in consistent but somewhat different
outcomes from that of cross-validation. For the criteria of
variance and correlation, the traditional regression analysis
shows better performance. Again, the count data model
performs the worst. The most obvious difference in the
model performance can be found from the measure of
coincidence. This index has been conceived to identify a
models ability to simulate patterns in the data. The
category-type models achieve remarkable performance in
this measure.
Thus, it can be summarized that the category-type
models show superior performance. The regression model
also generated an acceptable level of performance. Even

Transportation Letters: the International Journal of Transportation Research

2014

VOL

NO

Chang et al.

Comparative analysis of trip generation models

Table 10 The result of validation

Cross-validation
RMSE
% RMSE
Correlation coefficient
Coincidence ratio
Back-casting
RMSE
% RMSE
Correlation coefficient
Coincidence ratio

Regression

Tobit

Poisson

Ordered logit

Category

MCA

0.6787
62.16%
0.5405
0.7531

0.6808
62.34%
0.5402
0.7407

0.6865
62.86%
0.4899
0.5640

0.6812
62.38%
0.5471
0.7713

0.6423
58.82%
0.6107
0.7500

0.6466
59.22%
0.5975
0.7430

0.6126
55.04%
0.6241
0.6885

0.6391
57.42%
0.6037
0.6805

0.6548
58.84%
0.5389
0.5387

0.6499
58.39%
0.5764
0.6971

0.6456
58.01%
0.5761
0.8188

0.6478
58.21%
0.5704
0.8134

though this result does not directly mean that the betterperformance model should become the standard framework for trip generation forecasting, the conventional
models can be said to address a satisfactory trip-making
behavior.
It should also be noted that the more sophisticated
models have not shown better performance in terms of
validation, though their implementation is more onerous.
This does not mean that the models do not provide more
advanced frameworks. Indeed, the models have more
appealing theoretical bases compared to those of the
conventional approaches. It is difficult, however, to find
ways to improve the mechanism of trip-making behavior
of individuals against the traditional methods. It is one of
the causes for the disappointing performance of the
advanced models.

Conclusion
Trip generation forecasts the number of trips that begin
from or end in each travel analysis zone. This modeling is
the first phase of the four-step travel forecasting procedure. Traditionally, regression and category analyses have
been applied to this step and have generated an acceptable
level of performance from the perspective of planning.
However, structural limitations for each type have also
been criticized. The negative trip rate likelihood, the
continuous nature in trip rates, and the lack of incorporation of traveler behavior characteristics are the typical
problems for the regression model. The category analysis
has also suffered from the drawbacks of non-statistical
justification and cell-by-cell calculation of trip rates.
Several alternative approaches have been put forward,
but a systematic investigation into the performance
between the models had not sufficiently been performed.
This paper provides such an analysis, using the Seoul
metropolitan region as the study area. The household
travel diary data and REDIS were used for modeling. Six
kinds of models were estimated and validated. The results
show that the category-type models are superior in overall
performance.
The results of this kind of comparative study may be
specific to the datasets used. Some approaches may lead to
better replication of observed patterns; however, they may
not lead to better forecasts. Namely, the findings by this
study should not be understood to give a green light to the
traditional methodologies for being the standard framework

of trip generation modeling. However, this paper illustrates


that seemingly advanced methods, without refinement in
trip-making behavior mechanisms, do not necessarily
provide better performance in trip generation.

Acknowledgements
This work was supported by the BK 21 plus program of
the National Research Foundation of Korea.

References
Agyemang-Duah, K. and Hall, F. L. 1997. Spatial transferability of an
ordered response model of trip generation, Transp. Res. A Policy
Pract., 31, 389402.
Badoe, D. A. 2007. Forecasting travel demand with alternatively
structured models of trip frequency, Transp. Plann. Technol., 30,
455475.
Bastida, C. and Holguin-Veras, J. 2009. Freight generation models,
Transp. Res. Rec. J. Transp. Res. Board., 2097, 5161.
Barmby, T. and Doornik, J. 1989. Modelling trip frequency as a Poisson
variable, J. Transp. Econ. Policy, 309315.
Cambridge Systematics, Inc. 2010. Travel model validation and reasonableness checking manual, 2nd edn, Travel Model Improvement
Program, Federal Highway Administration, Washington DC, US.
Committee for Determination of the State of the Practice in Metropolitan
Area Travel Forecasting. 2007. Metropolitan travel forecasting
current practice and future direction, Special Report 288,
Transportation Research Board, Washington, DC.
Cotrus, A., Prashker, J. and Shiftan, Y. 2005. Spatial and temporal
transferability of trip generation demand models in Israel, J. Transp.
Stat., 8, 3756.
Daly, A. 1997. Improved methods for trip generation, Transportation
Planning Methods Volume 11, Proceedings of Seminar F held at PTRC
European Transport Forum, Brunel University, England, 207222.
Donnelly, R., Erhardt, G., Moeckel, R. and Davidson, W., 2010.
Advanced practices in travel forecasting, National Cooperative
Highway Research Program Synthesis 406, Transportation
Research Board, Washington DC, US.
Greene, W. H. 2000. Econometric analysis, 4th edn, London, Prentice
Hall International, Inc.
Jang, T. Y. 2005. Count data models for trip generation, J. Transp. Eng.,
131, 444450.
Kohavi, R. 1995. A study of cross-validation and bootstrap for accuracy
estimation and model selection, the 14th International Joint
Conference on Artificial Intelligence, Montreal, Quebec, Canada,
11371145.
Lim, K. and Srinivasan, S. 2011. Comparative analysis of alternate
econometric structures for trip generation models, Transp. Res.
Rec., 2254, 6878.
Ma, J. and Goulias, K. G. 1999. Application of Poisson regression models
to activity frequency analysis and prediction, Transp. Res. Rec. J.
Transp. Res. Board, 1676, 8694.
McNally, M. G. and Rindt, C. R. 2007. The activity-based approach, in
Handbook of transport modelling, (eds. D. A. Hensher and K. J.
Button), Amsterdam, Netherlands, Elsevier Science Ltd, pp. 5574.

Transportation Letters: the International Journal of Transportation Research

2014

VOL

NO

87

Chang et al.

Comparative analysis of trip generation models

Ortuzar, J. and Willumsen, L. G. 2011. Modelling transport, 4th edn,


Chichester, West Sussex, UK, Wiley.
Papacostas, C. S. and Prevedouros, P. D. 2001. Transportation engineering
and planning, Upper Saddle River, NJ, USA, Prentice Hall.
Picard, R. R. and Cook, R. D. 1984. Cross-validation of regression
models, J. Am. Stat. Assoc., 79, 575583.
Roorda, M. J., Paez, A., Morency, C., Mercado, R. and Farber, S. 2010.
Trip generation of vulnerable populations in three Canadian cities: a
spatial ordered probit approach, Transportation, 37, 525548.
Schmocker, J. D., Quddus, M. A., Noland, R. B. and Bell, M. G. H. 2005.
Estimating trip generation of elderly and disabled people: analysis of
London data, Transp. Res. Rec., 1924, 918.

88

Sheffi, Y. 1979. Estimating choice probabilities among nested alternatives,


Transp. Res. B Methodol., 13, 189205.
Stopher, P. and McDonald, K. 1983. Trip generation by crossclassification: an alternative methodology, Transp. Res. Rec., 944,
8491.
Train, K. E. 2009. Discrete choice methods with simulation, 2nd edn, New
York, NY, USA, Cambridge University Press.
Wallace, B., Mannering, F. and Rutherford, G. S. 1999. Evaluating
effects of transportation demand management strategies on trip
generation by using Poisson and negative binomial regression, Transp. Res. Rec. J. Transp. Res. Board, 1682, 70
77.

Transportation Letters: the International Journal of Transportation Research

2014

VOL

NO

You might also like