Вы находитесь на странице: 1из 11

Value of Dynamic Revenue-Maximizing Congestion

Pricing in a Highly Congested Corridor


John El Khoury 1 and F. Jordan Srour 2

Abstract: This paper examines the value of utilizing dynamic, revenue-maximizing, congestion pricing on a privately operated tolled route,
when the only alternative is a highly congested, public, free-access route. Three methods are specified by which revenue can be maximized—
Downloaded from ascelibrary.org by Cambridge University on 08/16/15. Copyright ASCE. For personal use only; all rights reserved.

through the selection of a fixed percent of users to set the toll price for, through the selection of a fixed level of service on the toll road, and
through the use of a nonlinear optimization model. These methods are tested on a case study involving a highly congested corridor.
In simulation, the corridor’s flow rates are observed following the posting of a toll, the toll is updated, and the new flow observed until
a subsequent round of revenue maximizing is applied to the new conditions. It was concluded that in a highly congested corridor a dynamic,
revenue-maximizing toll can finance the construction of new capacity without degrading social welfare. DOI: 10.1061/(ASCE)TE.1943-
5436.0000798. © 2015 American Society of Civil Engineers.
Author keywords: Congestion pricing; Dynamic toll pricing; Managed lanes; Corridor management; Revenue maximizing tolls.

Introduction recoup investment costs are gaining traction (Parry 2009; Zhao et al.
2010; Wang et al. 2011; Lindsey 2012; Chang and Hsueh 2006;
Developing countries tend to suffer from what is termed premature Hassan et al. 2013). This work introduces a strategy for dynami-
congestion (Gwilliam 2003). Premature congestion is characterized cally setting toll prices to maximize revenue. The proposed strategy
by rates of car ownership exceeding the predicted rates seen in is studied under three different cases—fixing the percentage of
developed nations based on the gross national income (GNI). roadway users allowed onto the toll road, fixing the toll road
For example, in the United Kingdom, the rate of car ownership is LOS, and dynamically optimizing toll prices to maximize revenue.
519 per 1,000 people, and the GNI per capita in international Before proceeding to describe the strategy for toll pricing, the
dollars based on purchasing power parity is 38,160; in contrast, authors review the most relevant literature, highlighting how this
car ownership in the developing country of Lebanon is 434 per work differs from or extends the current body of knowledge. Sub-
1,000 people, while the GNI is 17,400 (World Bank 2013). This sequently, the problem is defined and the nonlinear mathematical
leads to high levels of congestion, as the rate of infrastructure program for maximizing revenue is introduced. Next, the capabil-
development or government investment in transport alternatives, ities of the mathematical program applied to a case study on a
which requires tax revenue from income, lags behind vehicle highly congested roadway in Lebanon are demonstrated. Then,
ownership, which is, of course, fueled by the lack of alternatives. the results of this case study are presented. The paper concludes
In many locations, this condition is further heightened by land use with a discussion and directions for future research.
patterns, such as urbanization, that encourage people to commute
into cities for economic opportunities.
In addition to the economics driving premature congestion, de- Relevant Literature
veloping nations often suffer from a lack of coordination in policy
making and regulation (Gwilliam 2003). As a result, the private The literature on congestion pricing can be approximately divided
sector is left to provide the necessary services. Of course, no private into two main categories: fixed and dynamic pricing. In fixed pric-
entity will fund the construction of a highway without monetary ing schemes, one price is set for a given time period and users adapt
incentive. To that end, a strategy is presented for maximizing toll their behavior accordingly (Verhoef et al. 1996; Yang and Huang
revenues on a highly congested corridor in Lebanon. 2004). In dynamic pricing, however, the toll is set in reaction to or
The majority of road pricing studies focus on deriving toll prices anticipation of roadway conditions (Dong et al. 2011). Within both
to manage congestion with the goal of maintaining a good system fixed and dynamic tolling strategies, different goals can govern the
wide level of service (LOS) (Hearn and Ramana 1998; de Palma setting of the prices. Chief among these goals are the goal to man-
and Lindsey 2011). More recently, with the economic crises facing age utilization and LOS across the network and the goal to maxi-
many countries, models exploring the capability of setting tolls to mize revenue.
As the literature on tolling to manage the utilization and LOS
1
Dept. of Civil Engineering, Lebanese American Univ., P.O. Box 36, across the network is substantial and not the main aim of this paper,
Byblos, Lebanon (corresponding author). E-mail: john.khoury@lau.edu.lb the remainder of this review is focused on research addressing the
2
Dept. of IT and Operations Management, Lebanese American Univ., goal of revenue maximization. Looking at the policy underlying a
P.O. Box 13-5053, Chouran, Beirut 1102 2801, Lebanon. E-mail: jordan
revenue-maximizing strategy, Lindsey (2012) poses a series of four
.srour@lau.edu.lb
Note. This manuscript was submitted on February 2, 2015; approved on
questions about the social welfare implications of private tolling.
June 16, 2015; published online on August 13, 2015. Discussion period Specifically, Lindsey (2012) poses questions about the interplay
open until January 13, 2016; separate discussions must be submitted for of cost recovery and demand management, the interplay of pricing
individual papers. This paper is part of the Journal of Transportation En- and road capacity, the interplay of pricing and public transportation
gineering, © ASCE, ISSN 0733-947X/04015029(11)/$25.00. options, and the interplay of private investor/operator goals and

© ASCE 04015029-1 J. Transp. Eng.

J. Transp. Eng.
social welfare goals. Whereas the article provides an economic and in allowable flow by using a predictive scheme to set the willing-
policy view on the field of toll pricing, it does not guide the devel- ness to pay parameters, which, in turn, dictate the level of utiliza-
opment of models for optimizing toll revenue. tion on the toll road. The willingness to pay parameters, which are
Verhoef et al. (1996) were among the first to address the prob- iteratively estimated in a prediction module, serve to obfuscate
lem of second-best, one-route congestion pricing with an untolled the need to estimate delays on the roadway. This is in contrast
alternative and the goal of maximizing profit. The model of Verhoef to the presented work in which the delay savings toll users can ex-
et al. (1996) yields a fixed toll price that is premised on an invariant pect are explicitly estimated.
value of time (VOT) across roadway users. However, given a dis- Although the present work is closely related to that of Hassan
tribution of VOT in the population, the number of users would, et al. (2013), this work assumes a simpler roadway structure of one
instead, be a function of the toll price. Despite this assumption, tolled and one untolled route. As such, the literature on dynamic
Verhoef et al. (1996) demonstrate that using a revenue-maximizing pricing for high occupancy toll (HOT) lanes is relevant. Gardner
goal can have detrimental effects on the social welfare of road- et al. (2013) studied the effect of setting HOT prices to ensure
way users. full utilization of the HOT lanes. They found that ensuring full uti-
As a result of these poor social welfare implications, the trend in lization requires a dynamic toll pricing scheme that is dependent on
Downloaded from ascelibrary.org by Cambridge University on 08/16/15. Copyright ASCE. For personal use only; all rights reserved.

congestion pricing moved away from that of revenue maximization, the roadway user’s VOT. Extending this work to include the con-
until the financial crisis of 2008. At that time, policy makers began cept of revenue maximization, Jang et al. (2014) examined the
to realize that tolls may serve as a mechanism to recoup investment effect on revenue of varying the effective capacity of the HOT lane.
costs or encourage public-private partnerships (Parry 2009). Fan They found, empirically on a case in California, that revenue is
and Gurmu (2014) recognized that toll pricing and infrastructure maximized when the capacity of the toll lane is reduced by
investment are fundamentally intertwined, as they propose a ge- 300 vehicles=h. This study extends this work by dynamically man-
netic algorithm for the strategic problem of simultaneously decid- aging the effective capacity of the toll lane with the goal of revenue
ing on the extent of capacity expansion and the toll price. In maximization.
addition to simply recognizing that a single toll price can overcome The presented work represents a contribution to the literature
the cost of construction, other researchers have explored the inter- along four lines. First, the reconcile the model of Jang et al.
play of both usage and access tolling. Specifically, Wang et al. (2014) is reconciled with that of Hassan et al. (2013) to provide
(2011) demonstrate that it is always profitable to implement an ac- an objective function dependent only on estimations of delay and
cess fee scheme with potentially the same effect on social welfare a VOT function. Second, the versatility of the proposed objective
as a usage fee structure. Given the assumed technology and policy function in accommodating both reactive and anticipatory strate-
underlying their proposed system, these results are only instructive gies for pricing is shown. Third, the value of this optimization-
in the setting of fixed tolls, but do not address the issue of dynamic based strategy is demonstrated on a case study calibrated with
tolling. actual roadway traffic data and local, survey-based estimations
Also in the context of fixed tolls, but with the goal of implement- of the VOT. Finally, the authors present results indicating that in
ing congestion pricing without detrimental social effects, a series of a highly congested corridor dynamic revenue maximization can
articles examines what is known as Pareto-improving toll pricing. serve to alleviate congestion without degrading social welfare.
In these schemes, the basic premise is to set the toll price and allow
for revenue redistribution, via refunds, such that no user is worse
off with the congestion pricing in place. Specifically, Guo and Yang Problem Definition and Formulation
(2010), recognizing that the road users’ value of time influences
their travel behavior, prove that a Pareto-improving refunding Given a single route with both tolled and untolled options
scheme exists if and only if the total system monetary travel dis- (e.g., Verhoef et al. 1996), a dynamic toll price, πt , on the tolled
utility is reduced. While this result is theoretically appealing, it is road in time t is set in order to maximize revenue. The untolled or
practically impossible to administer refunds to roadway users general purpose (GP) lanes have a capacity of CGP and in any time
period t may carry a volume of V GP t ≤ C . Similarly, the parallel
GP
based on varying values of time. Therefore, Guo and Yang
(2010) also propose a single refund amount that is still Pareto im- toll road has a capacity of CT . However, to control the flow on
proving. Extending this work, Nie and Liu (2010) prove the exist- the toll road, a buffer is also specified, δ t , that reduces CT to an
effective capacity at time t: CET t ¼ C − δ t . The volume traveling
T
ence of a self-financing, Pareto-improving congestion pricing
scheme when the VOT function is concave. This is significant on the toll road in time t is V Tt . The total demand for the entire
as such a scheme would allow the tolling authority to keep, rather system at time t is Dt .
than redistribute, the net revenues. In keeping these funds, they may Given this setting, the expression for revenue in time t is
reinvest the money in the infrastructure. Although, as the proposed Rt ¼ πt V Tt . In turn, πt is a function of the amount of money that
toll scheme is static, its applicability to the setting examined in this users are willing to pay for an hour of time relative to the total
paper is limited. amount of time that they can save by taking the toll road. Thus,
As far as is known, there are only two papers that address dy- the expression for revenue can be expanded to Rt ¼ θt dt V Tt , where
namic road pricing for revenue maximization (Chang and Hsueh θt represents the value associated with an hour of time as set in time
2006; Hassan et al. 2013). The earlier of the two works addresses t to differentiate users, dt represents the delay that users of the toll
the need for profit maximization through the idea of a build- road can avoid (the difference between the delay on the GP lanes
operate-transfer arrangement for infrastructure development and that on the toll lane), and V Tt is as before.
(Chang and Hsueh 2006). Specifically, Chang and Hsueh (2006) The value of θt can be derived from the constraints regarding the
optimize toll revenues across a network of roads subject to equi- capacity and volume of the toll road and the users’ VOT. Specifi-
librium constraints. However, given the complexity of optimizing cally, following the analysis of both Jang et al. (2014) and Hassan
across multiple road links, they set the tolls based on an invariant et al. (2013), if the VOT is expressed as a cumulative distribution
VOT. Although they do present empirical results across various as- function, then it can be inverted to find the value of θt . Although
sumed levels of the VOT. In the more recent paper, Hassan et al. both previous papers assume a logit structure for the CDF of the
(2013) address the issue of optimizing revenue through variations VOT, this work leaves the function in a general form, FVOT ðθt Þ, and

© ASCE 04015029-2 J. Transp. Eng.

J. Transp. Eng.
shows how it can be manipulated to derive a nonlinear program for the calibration and use of this model on a highly congested stretch
maximizing revenue. of roadway in Lebanon are demonstrated.
Considering the constraints governing the effective capacity
and volume on the toll road, it is known, as in Jang et al. (2014),
that the effective capacity must be less than the total available Case Study
capacity on the toll road and greater than the demand wishing to
use the road. Symbolically, it can be written that CT − δ t ≥ CET While the simplicity of the model’s underlying structure of a single
t ≥
½1 − FVOT ðθt ÞDt . Manipulating these inequalities then 1 − ðCTt − route with tolled and untolled alternatives may seem unrealistic,
δ t Þ=Dt ≤ FVOT ðθt Þ. Now, in order to find θt , the inverse of given the topography of Lebanon’s coastline, it is actually reflective
FVOT ðθt Þ is taken, yielding the following toll price: of reality. Surrounded by mountains to the east and the Mediterra-
nean Sea to the west, Lebanon has a narrow coastal plain, which is
  225 km (139.8 mi) long. Beirut, the capital, is located in the center
CT − δ t
θt ≥ F−1
VOT 1 − ð1Þ of the north-south coastline. In 2000, the Greater Beirut area was
Dt
home to 1.3 million people (approximately one-third of the total
Downloaded from ascelibrary.org by Cambridge University on 08/16/15. Copyright ASCE. For personal use only; all rights reserved.

country’s population). The northern city of Tripoli is the second


Placing this expression into the previous revenue expression largest in the country with an estimated population of 450,000.
yields the following mathematical program for maximizing Being the major employment center, Beirut attracts trips from
revenue: the suburbs as well as the nearby cities. Only one major arterial
  (A1 Highway) runs along the narrow coastline and connects Tripoli
CT − δ t
maxF−1
VOT 1− dt V Tt such that 0 ≤ δ t < CT ð2Þ (and other major cities like Byblos and Jounieh) to Beirut. Besides
δt Dt the unorganized minibus public/private service (with no dedicated
stops and no dedicated times), there are no other public transpor-
Clearly, the form of the model depends on the functional form tation alternatives, forcing the majority of people to use private
that FVOT ðθt Þ takes. If a binary or mixed logit model is assumed as cars. As a result, weekday commute traffic flow into the city is
in Hassan et al. (2013) or Jang et al. (2014), respectively, then this highly directional, with peak flows inbound in the morning hours
mathematical program becomes a concave, nonlinear program. and outbound in the evening hours.
The method used to estimate the expected difference in delay The cross section of the A1 Highway varies from two lanes per
between the GP and the toll lanes will also affect the form of direction to four lanes per direction. Although the concept of lim-
the mathematical program. If the delay in period t is assumed to ited access does not work in Lebanon, where highway control/
equal the delay in period t − 1, then dt can be taken from the sensor enforcement and traffic management is lacking, the A1 Highway
data. However, doing so will potentially yield a monotonically de- represents the road with the highest level of mobility and least ac-
creasing function. In turn, given the myopic nature of always maxi- cess in the country, and as such, is labeled a principal arterial. To
mizing the revenue in each period, without considering the future perform this study, a model of the most congested section of the A1
effect of that decision, the optimal solution will always occur by Highway, spanning 8.9 km (5.56 mi), was the basis of the simu-
setting the buffer to zero and opening all of the capacity on the toll lation. As shown in Fig. 1, the project limits extend from south of
road. As a result, the overall revenue will not be maximized. Thus, the Tabarja interchange to south of the Zouk interchange. Five main
it is better to use a mechanism to predict the effect of the buffer interchanges exist within the project limits, between which volume
selected in the current period on the delay that will be seen in future counts and speed data were collected.
periods. To that end, a Bureau of Public Roads (BPR)-type estima-
tion of the delay on both the GP and toll lanes is utilized. The fol-
VOT Calibration
lowing expression is obtained for the delay that users can avoid
assuming a free-flow time of T 0 : Both the theoretical definition of the VOT and its actual measure are
 4  T 4  crucial to the cost analysis, evaluation of alternatives, development
V GP
t Vt of pricing policies, and the modeling of travel demand and behavior
dt ¼ 0.15T 0 − ð3Þ in any transport project. In the cornerstone theory behind the VOT,
CGP CT
Becker (1965) assumes that time can be transferred freely between
Recognizing that the BPR function does capture the effect of the work and leisure. Thus, when some savings are achieved in travel
δ t term but does not capture the overall propagation of delay in the time, it can be used for added productivity at work. For that reason,
system, a set of values, γ t , that will be negative in the start of a peak most investment projects aim at saving travel time by reducing de-
period and positive toward the end is added to the function. This lay, increasing the average travel speed, and/or providing shorter
strategy serves to correct the BPR’s tendency to overestimate the route options (Small 2012); besides, of course those projects
delay in the time steps leading to the peak in demand and under- enhancing highway safety.
estimate delay in the time steps following the peak in demand. Being one of the most widely discussed concepts in the trans-
Furthermore, if it is assumed that the roadway users will select portation economics literature, the VOT has been noted to vary by
to use the toll road based on the toll that is set, then V Tt can be trip purpose, time of day, congestion levels, income, gender, and
replaced by CT − δ t. As a result, the objective of Eq. (2) is ex- family status, among other factors (Small and Verhoef 2007). It is,
panded to become hence, understandable that the literature revealed wide variations in
measures of the perceived VOT. Since the article by Lave (1969),
   GP 4  T   
CT − δ t Vt C − δt 4 several studies show that commute trips average approximately half
max.15F−1
VOT 1 − T 0 − þ γ t of the wage rate (Small and Verhoef 2007; Hensher and Greene
δt Dt CGP CT
2011). Furthermore, it is well documented in the literature that
× ðCT − δ t Þ such that 0 ≤ δ t < CT ð4Þ VOT increases under congested versus free-flow conditions by ap-
proximately 25% and up to 55% (Abrantes and Wardman 2011).
In Eq. (4), regardless of the form of FVOT ðθt Þ, the model is non- This is because drivers usually exert extra effort to safely navigate
linear and concave over the range of allowable values for δt. Now through congestion. Additionally, the VOT increases with the

© ASCE 04015029-3 J. Transp. Eng.

J. Transp. Eng.
Downloaded from ascelibrary.org by Cambridge University on 08/16/15. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Project limits along A1 Highway (image © Google, Data SIO, NOAA, U.S. Navy, NGA, GEBCO)

income level, exhibiting an income elasticity of the VOT, an end of the survey, subjects were allowed to add their comments
average of 0.5 [Shires and de Jong 2009; Gunn 2001; Strategic related to the study and the proposed toll road. The survey could
Highway Research Program (SHRP 2) 2013] and was measured be filled out in either English or Arabic.
by Abrantes and Wardman (2011) to reach 0.9. In fact, Borjesson As the primary goal of the survey is to estimate VOT, which is
et al. (2012) find it tending to zero for low incomes and more than continuous in nature, the sample size calculations were made ac-
1.0 for higher incomes. cordingly. Specifically, with a desired confidence level of 95%, an
Ultimately these factors can be captured in a cumulative distri- acceptable level of error of 250 Lebanese Lira (LL), and an esti-
bution function, FVOT ðθt Þ. This function represents the percentage mated standard deviation of 4,000LL (US$2.66), the target sample
of the population that is willing to pay more than θt to save 1 h of size was 983. This sample size represents approximately 2% of the
time. In this case study, a survey was conducted to derive the target population—the users of the existing A1 Highway between
functional form most descriptive of this region and user group the Tabarja and Nahr El Kalb interchanges where commuter traffic
of interest. The survey follows a stated preference design. The is estimated at 53,000 vehicles per day per direction (Council for
stated preference design serves to capture the users’ perceived Development and Reconstruction 2011).
sense of time; the perceived sense of time is particularly relevant The survey was administered both online and as an intercept
to toll pricing as it is a user’s perception that will ultimately drive survey at restaurants and service stations along the A1 Highway
their decision to use a toll lane or not. within the study limits. The online survey was disseminated as
The first part of the survey (Q1 through Q6) served to determine a link via e-mail to user lists at the Lebanese American University,
the respondent’s use pattern and experience on the roadway seg- Byblos Bank, and Kunhadi (a roadway safety awareness organiza-
ment in question. The second part (Q7 through Q11) focused on tion in Lebanon). The link to the survey was also published via the
assessing the respondent’s willingness to pay various fee levels Web page of a well-known TV anchor in Lebanon (on the Lebanese
across various potential time savings; a question was also asked Broadcasting Company Incorporated website). Regarding the sur-
to determine the respondent’s familiarity with the tolling concept. veys conducted via intercept, a typical interview lasted anywhere
Finally, the last set of questions (Q12 through Q14) served to solicit from 2 to 8 min, depending on the availability and the attentiveness
the respondent’s demographics (age, income, and gender). At the of the respondents. The interviewers targeted mainly coffee shops,

© ASCE 04015029-4 J. Transp. Eng.

J. Transp. Eng.
grocery stores, restaurants, and gas stations within the project Thus, roadway users in Lebanon have a VOT that is almost 0.6
limits. of their wage rate, which is consistent with values in the literature.
The online survey was closed when the total number of re- A graphical depiction of the empirical results and the fitted distri-
sponses from both the online and intercept surveys reached 1,219. bution can be seen in Fig. 2.
However, of those, 13 were left blank and were thus removed. Of
the remaining 1,206, 70 respondents claimed that they never trav- Traffic Flow Calibration
eled the stretch of roadway in question. As such, they were re-
moved from the sample. A further 12 respondents did not provide Traffic data collection techniques primarily fall under three major
any data regarding their willingness to pay for various levels of categories: site data, floating car data (FCD), and wide-area data.
time savings and were also removed. Finally, given the use of social Each of these methods provides different traffic data, accuracy, and
media as a means for survey dissemination, the responses showed network coverage (Antoniou et al. 2008).
a disproportionate number of young (18–24) respondents (63%). Given the lack of an organized traffic management center
As a result, a stratification process is utilized. (TMC), the absence of government and local agency support, vir-
Recognizing that the surveys administered via intercept were tually nonexistent project specific funds, and no permanent vehicle
Downloaded from ascelibrary.org by Cambridge University on 08/16/15. Copyright ASCE. For personal use only; all rights reserved.

more reflective of the age distribution of roadway users, the relative detection or count stations along the A1 Highway, the authors relied
proportions of those respondents were utilized to formulate a sam- on site data and FCD collected by a team of students. For the former,
pling procedure among the remaining responses. Specifically, the manual counters and digital cameras were used at key locations
required sample strata sizes were identified, as follows: 39% age along the corridor—mainly between the main interchanges—to
18–24, 28% age 25–34, 28% age 35–55, and 5% older than 55. collect peak hour traffic count data. Digital camera recordings were
To achieve this desired split, this study targeted, for culling, a group later analyzed to obtain traffic counts at those locations. Peak hour
of 286 respondents (25%) who reported being between 18 and traffic counts were collected for 2-h periods in both peak periods
24 years old and traveled the roadway less than two times per week (7:00 to 9:00 a.m. and 4:00 to 6:00 p.m.). At each location, peak
or only 2–3 times per week. These respondents were likely stu- hour traffic was counted for two days (some combination of
dents, given the used sampling strategy. Thus, from this group, Tuesday, Wednesday, and Thursday) during the period from
152 responses were randomly selected to be removed; all responses September 2014 to December 2014. Counts were conducted outside
from the other age groups were included for analysis. While only official or school holidays and during fair weather to eliminate
464 of the responses reflected regular travel on the roadway seg- the effects of rainy days/storms, holidays, and weekend return/
ment of interest, nearly all respondents were still able to provide departure occurring on Mondays and Fridays. Peak hour data
estimates on their willingness to pay for different levels of travel was also reported by direction and in 15-min increments [to obtain
time savings. This left a total of 972 responses for analysis. the peak hour factor (PHF)]. Midday and off-peak hour traffic counts
The information on the respondents’ stated willingness to pay were also collected. Those counts were conducted for only one day
relative to their expected time savings was particularly robust, as during the week, which was based on the availability of the student
each respondent provided information on their willingness to pay assistants to be onsite. The accumulation of the hourly data: peak, mid-
for less than 10 min of savings, between 11 and 20 min, between 21 day, and off-peak, helped provide a daily distribution of traffic counts.
and 30 min, between 31 and 45 min, between 46 and 60 min, and Truck counts/percentages were collected during midday and off-peak
over an hour. The increments of 10 min at the low end of the scale hours only, as trucks are not allowed to operate during the commute
were smaller than at the high end (increments of 15 min) in order to peak hours per government regulations. The location of the digital cam-
more accurately capture the respondents’ perception of time. For era was selected to be at a vantage point where a long stretch of the road
example, it is much harder for a person to rationally assign a mon- was captured. This allowed for the capture of the traffic densities in
etary value to an amount of time less than 10 min, but much easier each of the hours analyzed—and specifically the jam densities.
to assign a value to an amount of time greater than a half hour. the In addition to traffic count data, speed data was collected using
respondents’ estimated willingness to pay for each time length into two methods. The first method was based on the video recordings at
estimates of their value associated with 1 h was converted. For
example, if a respondent stated that they would pay 1,000LL
(US$0.66) to save less than 10 min on the road, this is equivalent
to being willing to pay 6,000LL (US$4.00) for 1 h.
Once the various values of time were assembled, a histogram
reflecting the empirical probability distribution function was con-
structed over price ranges of 1,000LL (US$0.66). From this prob-
ability density function (PDF), a cumulative distribution function
(CDF) was constructed and a distribution was selected for fitting.
The Weibull distribution was chosen to fit the data, as it showed the
best fit along with a functional form that allowed for scaling of
the currency term. Given the relatively large denominations seen
in the Lebanese Lira, with 5,000LL (US$3.33) serving as a com-
monly used bill, the presence of a scaling term in the CDF was
important. The exact expression of the VOT, obtained by fitting
the Weibull CDF was

FVOT ðθt Þ ¼ 1 − e−ðθt =7,309Þ


1.3
ð5Þ

On average, roadway users value their time at almost


8,000LL=h (US$5.33=h). This is consistent with the average wage
Fig. 2. Empirical PDF and CDF with fitted Weibull CDF
rate of almost 13,500LL=h (US$9=h), as reported on the survey.

© ASCE 04015029-5 J. Transp. Eng.

J. Transp. Eng.
the key locations along the corridor. More than one digital recorder • Capturing the prevailing traffic patterns within the corridor;
was used simultaneously, which allowed tracking signature ve- • Estimating flows that will use the toll lane;
hicles as they crossed the key locations. Signature vehicles were • Replicating geometric features of the GP and the toll lanes;
randomly selected vehicles having noticeable color, make, or fea- • Verifying traffic flows and speeds at various sections within the
tures. The digital cameras’ clocks were synched to allow time GP and toll corridors; and
tracking of such signature vehicles. Noting the time when a signa- • Testing the model sensitivity to varying parametric and geo-
ture vehicle passed each of the marked locations, and knowing the metric changes.
constant distance between the marked locations, the vehicle run- A variety of microscopic traffic models are capable of modeling
ning speed was computed using the space-mean-speed equation, freeway operations (including lane changes and weaving), static
Ū s ¼ l=t̄, where Ū s is the space mean speed, l is the length of and dynamic route assignment, user group/vehicle classification
section analyzed, and t̄ is the average travel time needed to cross features, and toll lane operation. For practical reasons, the selected
the given length (Ross 1991). model should be able to present the results of the simulation in such
In the second method, the FCD was used to collect average a way that descriptive statistics can be easily generated (Depart-
operating speeds within the corridor. At random times during week- ment of Transportation, Federal Highway Administration 2002).
Downloaded from ascelibrary.org by Cambridge University on 08/16/15. Copyright ASCE. For personal use only; all rights reserved.

days, students, and colleagues timed their trips between marked The traffic software integrated system (TSIS)/corridor simulation
locations within the 8.9 km section of the corridor. Again, using (CORSIM) tool, which was readily available, was used to assess
the equation above, the speed was computed per the analyzed hour. traffic operations within the freeway corridor. Using CORSIM,
Timed trips were conducted by multiple roadway users and during the geometry and flow of the existing A1 Highway corridor were
various morning and evening hours in both directions of the A1 modeled.
Highway within the project limits. At first, a working model of the existing corridor, termed the
Finally, 24-h traffic count and speed data for the southbound base model, was compiled using default CORSIM values. However,
direction were marked at the key locations within the project study such values cannot be expected to capture operations along the A1
limits. Corridor average speed and flow curves were computed and Highway, whose human and physical attributes are totally different
are plotted in Fig. 3. The figure demonstrates the well-defined peak than those seen in the U.S. Thus, the calibration of the model was a
in the southbound direction (into Beirut) during the morning peak crucial step in replicating local conditions, including driver behav-
period, which lasts from 6:30 to 9:30 a.m. Traffic jams in the south- ior, lane capacities, and speed/flow profiles (Hourdakis et al. 2003).
bound direction during morning hours (similarly in the northbound The objective of the calibration process, which was entirely based
direction during p.m. hours) stretch for kilometers and last for on Chapter 5 of Calibration of the Traffic Analysis Toolbox Volume
hours, causing enormous pollution problems, driver frustration IV: Guidelines for Applying CORSIM Microsimulation Modeling
and stress, lost productivity and a significant waste of fuel. The Software, was to find the set of parameter values for the model that
flow, speed, and travel time data was used to calibrate the simula- best reproduces local traffic conditions (Holm et al. 2007). The first
tion model as discussed in next section. step included calibrating the capacity of the GP lanes at key bottle-
necks within the corridor. Three major bottlenecks were identified
within the 8.9 km section of the A1 Highway, the first at the lane
CORSIM Calibration Process
reduction at Ghazir bridge (reduction from three to two lanes), the
A fundamental aspect of the toll lane operations is the determina- second at the heavy merge section from the high-volume Jounieh
tion of which toll price and corresponding traffic volume are ap- on-ramp with no acceleration lane, and the third at the diverge to
propriate at any given time, and at what point it is necessary to Kaslik interchange which eliminates a freeway lane with the off-
change the tolling scheme to maintain high efficiency within the ramp (reduction from 3 to 2 lanes). Because the capacity of the
toll lane. Because the toll lanes do not actually exist, a microscopic downstream bottleneck was controlling the two upstream bottle-
simulation model was used to assess operations on the existing GP necks, it was used for calibration. Additionally, a 10 × 10 O-D
lanes and the proposed toll lane. The modeling requirements for the matrix was developed for the base model to better match flows
toll lane analysis included: at the off-ramps throughout the corridor.
Three global parameters were systemically adjusted to calibrate
the model output. Those include the car following sensitivity
factor, the Pitt car following constant, and the percent of drivers
yielding the right-of-way (ROW) for lane changing drivers attempt-
ing to merge ahead. Based on Zhang and Holm (2004), those
parameters were most influential in modifying freeway capacity.
The car following sensitivity factor was adjusted from a default
mean of 0.8 s to a mean of 1.05 s. It represents the primary factor
in calculating the desired time headway (in seconds) between a
leader-follower pair. Increasing this value reduces freeway capac-
ity, which is commensurate with reality of the A1 Highway oper-
ations. The Pitt car following constant was also changed to 1.5 m
(5 ft) instead of a default value of 3.05 m (10 ft) to better represent
the local driving habits (drivers following very closely). Finally, the
percent of drivers yielding the ROW parameter was changed to 60%
instead of the 20% default value, as drivers tend to be more aggres-
sive in lane changing forcing other drivers to yield. There are other
parameters related to the lane changing logic within CORSIM;
however, altering these parameters is typically not worthwhile be-
Fig. 3. Southbound A1 Highway—corridor average hourly speed and
cause they are not very sensitive to aggregate measures (i.e., 15-min
flow
averages) of throughput (Zhang and Holm 2004). To determine the

© ASCE 04015029-6 J. Transp. Eng.

J. Transp. Eng.
number of model runs needed, the sample size equation, n ¼ 5. Optimize the buffer, δ t by solving Eq. (4) with the volume on
½ðzα=2 Þ2 × σ2 =E2 was used, where zα=2 is 1.96 for 95% confidence the GP lanes estimated according to the forecast model in
level, the sample variance, σ, is conservatively taken to be 1=2, and operation and the demand of the system assumed from historic
the desired margin of error, E, is assumed to be 10%, knowing that data;
“due to the stochastic nature of CORSIM results from individual 6. Based on the buffer set in Step 5, calculate πt using Steps 1 and
runs can vary by 25% and higher standard deviations may be ex- 3; and
pected for facilities operating at or near capacity” (Holm et al. 7. If t < T, increment t and return to Step 2. Otherwise, terminate.
2007). Although the required number of runs was found to be As the simulation structure recalculates the toll at every time step,
29, the base model was run in excess of 50 times before adopting the length of the time step should be chosen in compliance with any
the final values of the three parameters. regulations on the frequency of toll price changes. In the analyzed
The base model was calibrated for three main parameters, which case, the authors selected 15 min as a default time step. The simu-
are flow (vehicles=h), speed (km=h), and average travel time during lation is run incrementally to avoid losing the dynamics from pre-
the peak hour. The results of the calibration are shown in Table 1. vious time steps. That is, every time the simulation is run for a
Field measured flow, speed, and travel time data were compared specific time step, the previous time steps were run again to account
Downloaded from ascelibrary.org by Cambridge University on 08/16/15. Copyright ASCE. For personal use only; all rights reserved.

with the base model values at three bottleneck locations within for the cumulative effects of operations within the corridor.
the corridor limits. The overall corridor average travel time per ve- The presented simulation is also dependent on a forecast of the
hicle during the peak hour (7:00 to 8:00 a.m.) using the model was volume on the GP lanes in Step 5. To study the effect of this fore-
approximately 28.4 min, which matches the timed measurements cast, three different forecasting methods were examined. The first,
through the corridor (average of 29 min). The measured free-flow termed reactive, in which the volume in time t was set to the vol-
travel time is approximately 7 min compared with 6.9 min output ume in time t − 1. The second, termed anticipatory, in which the
using the base model. volume on the GP lanes in time t was set to be the demand antici-
Because the calibration targets were met, the model represents a pated in time t minus the observed volume on the toll road in time
close match to current field conditions. The base model was then t − 1. Finally, the third, termed exact, assumes that the volume on
revised to include the toll lanes (one lane per direction) parallel to the GP lanes in time t will be the anticipated demand minus the
the current GP lanes. The toll lanes ingress and egress points were exact volume on the toll road for which the toll price was set. That
t ¼ Dt − ðC − δ t Þ. In this case, the volume
is, in the third case, V GP T
assumed at locations where ROW/space permits gantry locations.
on the GP lanes becomes dependent on the decision variable δ t .
Finally, the capacity of the corridor was set at 4,000 vehicles=h
Finally, use of the BPR function in Eq. (4) requires the addition
on the GP lanes based on observed maximum flow just downstream
of the terms γ t to correct for the influence of delay propagation.
of the south-most bottleneck (between the Kaslik and Zouk
These terms were calibrated by studying the difference between
interchanges). This value defines the corridor’s maximum flow
the expected and actual delay witnessed at various levels of δ t fixed
under prevailing conditions. As for the toll lanes, a maximum
across the peak period. These differences were then averaged in
capacity of 1,800 vehicles=h was assumed to maintain good
each time period and the average was used as the γ t value. This
LOS at all times (Department of Transportation, Federal Highway
yielded values for the adjustor term, γ t , ranging from −0.1 to
Administration 2002). Finally, the corridor demands were projected −0.8 in the time steps leading up to the peak in demand and values
10 years into the future with average annual growth of 4%, given ranging from 0.2 to 1.1 in the periods following the peak in de-
the widespread housing developments and the influx from neigh- mand. Using the mean absolute percent error (MAPE) between
boring countries due to security issues, leading to an overall aver- the actual delay (measured using the CORSIM model) and the fore-
age increase of 43% over current traffic flows, which would also casted delay using the BPR function, it was found that, on average,
capture induced demand on this corridor. the addition of the γ t values improved the delay estimation by 44%
for the dynamic forecasting methods. On average, the delay MAPE
Simulation Structure was reduced from 57% difference by only using the BPR function
to 32% difference when adding the γ t values.
The simulation structure that was used is similar to that of Gardner Once the volume on the GP lanes is forecast and the BPR es-
et al. (2013). Specifically, a series of linked spreadsheets with calls timation of delay adjusted, the nonlinear program, represented in
to the CORSIM traffic simulator were used to implement the fol- Eq. (4), is rapidly solved by any commercially available, off-
lowing seven steps: the-shelf solver. In the presented case, it was found that the
1. Initialize: Set t ¼ 0 and the volumes on the toll and the GP built-in nonlinear solver of Excel could capably solve the model.
lanes according to an initial buffer of 0 (δ 0 ¼ 0). Set the toll
price to 500LL (US$0.33) as there will not be any delay ex-
perienced in the first period of simulation; Results
2. Run the roadway simulation for the length of one time step;
3. Calculate the revenue found in time t based on the actual vo- For the sake of comparison, the authors define two models: the base
lumes recorded on the toll road in time t; model and the untolled lane model. The base model represents
4. Increment the time step, set t ¼ t þ 1; existing conditions while the untolled lane model assumes the

Table 1. Measured versus Modeled Flow, Speed, and Travel Time at Three Bottlenecks within the Study Corridor
Flow (vehicles=h) Average speed (km=h) Average travel time (min)
Cumulative
Bottleneck location distance (km) Field Model Difference (%) Field Model Difference (%) Field Model Difference (%)
Maameltein bridge 1.9 3,560 3,725 5 10 13 30 11 9 −12
Jounieh on-ramp 4.2 3,970 4,025 1 10 12 20 23 22 −6
Kaslik off-ramp 7.1 4,165 4,055 −3 34 38 12 27 26 −5

© ASCE 04015029-7 J. Transp. Eng.

J. Transp. Eng.
Table 2. Summary of Results across Scenarios
Network throughput Average speed Total delay Average delay Average travel Total revenue
Analyzed scenario (vehicles=km) (km=h) (vehicles=h) (min=km) time (min) (US$)
Base model 115,071 31.4 2,205.5 1.15 17.8 0
Untolled lane 152,998 56.7 877.2 0.35 11.4 0
Fixed price point, US$6 142,377 42.2 1,657.6 0.70 12.4 8,247
Fixed buffer, 700 vehicles=h 144,285 40.3 1,839.7 0.76 13.4 9,771
Reactive 136,121 37.9 1,920.3 0.85 14.4 8,384
Anticipatory 139,182 38.3 1,934.3 0.83 15.4 12,833
Exact 130,364 35.4 2,065.1 0.95 16.4 16,091

addition of one untolled lane per direction with the same capacity the percentage split between the toll lane and GP lane users fixed.
Downloaded from ascelibrary.org by Cambridge University on 08/16/15. Copyright ASCE. For personal use only; all rights reserved.

of the envisioned toll lanes. For all scenarios, only the morning Given the VOT in the project population, the total revenue curve is
peak period (3 h from 6:30 to 9:30 a.m.) is analyzed as the after- found to be concave relative to θt with the maximum revenue of US
noon peak behaves similarly, but in the opposite direction. Besides, $8,247 for the morning peak period occurring at a price point of
the a.m. peak is a more defined peak than the p.m. peak, which is 9,000LL (US$6). Fig. 4(a) depicts these results graphically. In con-
observed to have a more prolonged duration and hence a little trast, a maximum revenue can be gained, for the morning peak
milder. It was found that the base model, averaged across the period, of US$9,771 by fixing the buffer on the toll road, δ t , to
3 h of the modeled period, allows a throughput in vehicle kilo- 700 vehicles=h. This level reduces the effective capacity of the toll
meters of 115,071, whereas the untolled lane model allows road to 1,100 vehicles=h. Thus, the tolling authority is better off
152,998 vehicles=km; an improvement of 33%. The average speed selecting a LOS for the toll road, and fixing that LOS in terms of
across the 3-h period in the base model is 31.4 km=h, whereas that a buffer, than selecting a percent of drivers to defer to the toll road.
in the untolled lane model is 56.7 km=h; an improvement of 81%. This result is consistent with that of Jang et al. (2014). Fig. 4(b)
Total delay over the three hours also improves by 60% with a delay depicts these results graphically.
of 2,205.5 vehicles=h in the base model versus 887.2 vehicles=h in In contrast to the scenarios above where the price point or buffer
the untolled lane model. Finally, the average travel time, across the is fixed, there are three scenarios where δt is the decision variable of
3-h period, on the main GP lanes (not the additional lanes) is a revenue-maximizing optimization. Specifically, the objective
17.8 min in the base model versus 11.4 min in the untolled lane function in Eq. (4) was used to set the toll price in each time step.
model. These results are summarized with the full set of results The element differentiating these three scenarios is the method by
in Table 2. The base model shows the nature of highly congested which the volume on the GP lanes is estimated. Figs. 5(a–c) illus-
GP lanes given the demand projected for the next 10 years. The trate how δ t changes across the time steps of the simulation relative
addition of the untolled lane provides significant operational ben- to the revenue realized in each of those time steps for the reactive,
efits by reducing delay and increasing throughput and average anticipatory, and exact methods of forecasting the volume on the
speed. Even with the addition of one toll lane per direction and GP lanes.
the shown operational improvements, it is anticipated that the min- From these results, it can be seen that forecasting the volume on
imal-induced demand will be generated due to the highly congested the GP lanes by using volume data from the previous time step
corridor. Despite the lane addition, the corridor capacity remains (reactive), puts a lag in setting the buffer that reduces the revenue
below demand levels. Of course, the addition of the lanes requires seen in the early periods of the simulation. Once the peak demand
substantial capital investment that may only be possible with toll- hits approximately 8:00 a.m., the buffer begins to decrease with a
ing. As such, the attention is now turned to the primary results of steady increase in revenue. The total revenue in the reactive case is
this study. US$8,384, which is less than that found in using a fixed buffer
First, the results associated with keeping the price point, θt , that strategy. In contrast, using information about the expected demand
differentiates users fixed is explored. This has the effect of keeping in the next time step coupled with actual toll lane volume data from

(a) (b)

Fig. 4. (a) Total revenue curve relative to different fixed price points, θt ; (b) total revenue curve relative to different fixed buffers, δt

© ASCE 04015029-8 J. Transp. Eng.

J. Transp. Eng.
Downloaded from ascelibrary.org by Cambridge University on 08/16/15. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

(c)

Fig. 5. (a) Changes in δt and total revenue across the time periods of simulation for the reactive method; (b) changes in δt and total revenue across the
time periods of simulation for the anticipatory method; (c) changes in δ t and total revenue across the time periods of simulation for the exact method

the previous step has the effect of holding the buffer within a range using a fixed price point of US$6, can bring total revenue of US
between 750 and 1,000 vehicles=h. This results in total revenue of $8,247 in the morning peak period. Of course, if a private toll author-
US$12,833. The fact that this range is consistent with the revenue- ity wishes to maximize revenue at the expense of the traveling pub-
maximizing, fixed buffer level of 700 is not surprising. In the exact lic, then a total revenue of US$16,091 can be made in the morning
case, where the volume on the GP lanes is predicted from expected peak while travelers will experience a total of 2,065.1 vehicles=h of
demands and assumed toll lane volumes, the buffer begins high in delay and an average speed of 35.4 km=h—this case represents only
order to increase delay on the GP lanes while capitalizing on a toll a 6% decrease in total delay compared with the base model, but im-
set to attract elite drivers; as the delay on the GP lanes increases proves total corridor throughput, over the case with no additional
through the peak period, the buffer is reduced to capitalize on roadway (the base model), by 13%, which is a measurable improve-
the volume of drivers wishing to access the toll road. As a result, ment. The willingness of users to pay for a toll under the exact strat-
in the exact case, the total revenue seen in the peak period is US egy regime could be questioned when the level of delay in vehicle
$16,091. These results imply that if the data regarding roadway hours is comparable to the base case. However, the level of delay
volumes can be accurately forecast, the dynamic optimization strat- specified here is total delay—which is mostly contributed to by
egy can serve to increase revenue substantially. Furthermore, with the delay seen on the GP; toll road users see negligible delay.
revenues of US$16,091 per morning peak, revenues of US$126 mil- In contrast to the dynamic strategies, the reason behind the
lion can potentially be gained over the span of 15 years assuming strong performance of the fixed price point strategy in terms of
that the afternoon peak would be tolled similarly. throughput is the fact that revenue maximization occurs when delay
Of course, the drive to maximize revenues should not be done at savings are at their highest on the toll road. Thus, in keeping the toll
the expense of the traveling public. Table 2 summarizes a variety of road for elite users both the willingness to pay and the delay savings
social welfare metrics across all scenarios. In these results it is remain high.
found that opening additional capacity in a highly congested cor-
ridor is the best option (the untolled lane case), as it allows for the
maximum throughput. However, without the capability to finance Discussion
such capacity a toll must be instituted. To that end, the tolling strat-
egies show that degrading the average speed, over the untolled lane In this paper, the authors demonstrated that a dynamic toll strat-
case, to 42.3 km=h with a total delay of 1,657.6 vehicles=h, by egy designed to maximize revenue can effectively allow the tolls

© ASCE 04015029-9 J. Transp. Eng.

J. Transp. Eng.
collected to pay for the infrastructure investment. However, if a swarm to the coast. It is well known that the purpose of the trip can
proper LOS is desired on the GP lanes as well, then a simpler toll- influence the VOT—as such, tolling during these leisure-related
ing strategy that fixes the percent of users to divert to the toll road peaks may yield very different results.
relative to a price point associated with their VOT is the best strat-
egy. In such a strategy, the government would need to subsidize
the construction of the roadway in order to recover the costs. Never- Acknowledgments
theless, understanding the effect of revenue-maximizing, dynamic
tolls on both revenue streams and LOS is important for policy The authors gratefully acknowledge the Lebanese American
makers. University’s School of Engineering Research Council for support
Moreover, when selecting a dynamic tolling strategy from the of this research.
three strategies presented in this paper, the tolling authority should
carefully consider the information requirements that each strategy
bears. The Reactive strategy requires the tolling authority to have References
the capacity to capture, in real time, the volume of users on both the Abrantes, P. A., and Wardman, M. R. (2011). “Meta-analysis of UK values
Downloaded from ascelibrary.org by Cambridge University on 08/16/15. Copyright ASCE. For personal use only; all rights reserved.

tolled and GP lanes. The anticipatory strategy requires both the of travel time: An update.” Transp. Res. A, 45(1), 1–17.
capability to capture roadway conditions and to mathematically Antoniou, C., Balakrishna, R., and Koutsopoulos, H. (2008). “Emerging
forecast the number of users on the GP lanes in the next time step. data collection technologies and their impact on traffic management ap-
Finally, the exact case requires the tolling authority to capture road- plications.” Proc., 10th Int. Conf. on Application of Advanced Technol-
way conditions and accurately forecast demand on both the GP and ogies in Transportation, National Technical University of Athens,
toll lanes. If the capability of forecasting demand accurately is de- Athens, Greece.
Becker, G. S. (1965). “A theory of the allocation of time.” Econ. J., 75(299),
graded, then the level of revenue from dynamic tolling will also
493–517.
likely be degraded. Furthermore, if implementing the appropriate Borjesson, M., Fosgerau, M., and Algers, S. (2012). “On the income elas-
hardware and software solutions to institute dynamic tolling is too ticity of the value of travel time.” Transp. Res. A, 46(2), 368–377.
costly, then the benefit of the exact strategy revenue will not be Chang, M.-S., and Hsueh, C.-F. (2006). “A dynamic road pricing model for
realized as profits. freeway electronic toll collection systems under build-operate–transfer
In addition to the investment in real-time roadway monitoring arrangements.” Transp. Plann. Technol., 29(2), 91–104.
technology, proper toll management implies posting the travel Council for Development and Reconstruction. (2011). “Traffic count for
times and toll prices well in advance of the toll road entrance. Such upgrading of A1 Highway between Nahr El Kalb and Tabarja (July)”.
technology is widely applied and would be feasible to install. de Palma, A., and Lindsey, R. (2011). “Traffic congestion pricing method-
Beyond the simple posting of the travel times and toll prices, ologies and technologies.” Transp. Res. Part C Emerg. Technol., 19(6),
1377–1399.
whether drivers would behave rationally relative to this information
Dong, J., Mahmassani, H. S., Erdoğan, S., and Lu, C.-C. (2011). “State-
could be speculated. In the fixed price or fixed buffer cases, having dependent pricing for real-time freeway management: Anticipatory ver-
drivers select the toll lane in numbers less than forecast would sus reactive strategies.” Transp. Res. Part C Emerg. Technol., 19(4),
cause an increase in congestion on the GP lanes and a reduction 644–657.
in revenues on the toll lane. If, instead, drivers select the toll lane Fan, W., and Gurmu, Z. (2014). “Combined decision making of congestion
in numbers greater than forecast, then the toll road might experi- pricing and capacity expansion: Genetic algorithm approach.” J.
ence some congestion and higher revenue—although with higher Transp. Eng., 10.1061/(ASCE)TE.1943-5436.0000695, 04014031.
congestion on the toll lane, it is expected that the more price Gardner, L. M., Bar-Gera, H., and Boyles, S. D. (2013). “Development and
sensitive users would revert to the GP lanes until the proposed comparison of choice models and tolling schemes for high-occupancy/
toll (hot) facilities.” Transp. Res. Part B Method., 55, 142–153.
equilibrium was reached. The effect of misjudging the driver
Gunn, H. F. (2001). “Spatial and temporal transferability of relationships
behavior relative to the posted travel times and toll prices is miti- between travel demand, trip cost and travel time.” Transp. Res. Part E
gated by the dynamic strategies. In these strategies, the continual Logist. Transp. Rev., 37(2–3), 163–189.
monitoring of roadway conditions allows for the response to un- Guo, X., and Yang, H. (2010). “Pareto-improving congestion pricing and
expected driver behavior in real time through the increase or de- revenue refunding with multiple user classes.” Transp. Res. Part B
crease of toll prices. Method., 44(8–9), 972–982.
This study is not without its limitations. First, the simulation Gwilliam, K. (2003). “Urban transport in developing countries.” Transp.
assumes that the toll lanes exist and that the traveling public will Rev., 23(2), 197–216.
choose to use the toll lane relative to the derived VOT. Of course, Hassan, A., Abdelghany, K., and Semple, J. (2013). “Dynamic road pricing
the decision to use the toll lanes may be governed by other factors for revenue maximization.” Transportation Research Record, 2345,
Transportation Research Board, Washington, DC, 100–108.
than simply VOT—users may wish to pass by certain stores or visit
Hearn, D., and Ramana, M. (1998). “Solving congestion toll pricing mod-
specific town centers on their way to and from work, thus forcing els.” Equilibrium and advanced transportation modelling, P. Marcotte
them to use the GP lanes. Second, the toll prices were set based on and S. Nguyen, eds., Springer, New York, 109–124.
the expected delay that toll users can avoid. Estimating this delay Hensher, D. A., and Greene, W. H. (2011). “Valuation of travel time savings
was done by using the BPR function, a function with known lim- in WTP and preference space in the presence of taste and scale hetero-
itations. Chief among those limitations is the fact that in the BPR geneity.” J. Transp. Econ. Policy, 45(3), 505–525.
function, volume drives the expected delay in any one period. Holm, P., Tomich, D., Sloboden, J., and Lowrance, C. (2007). “Traffic analy-
However, in reality, delay propagates thus making the actual delay sis toolbox volume IV: Guidelines for applying CORSIM microsimula-
lower initially and higher in later periods. While an attempt was tion modeling software.” 〈http://ops:fhwa:dot:gov/trafficanalysistools/
tatvol4/index:htm〉 (Oct. 2014).
made to correct for this shortcoming by the use of adjustor terms,
Hourdakis, J., Michalopoulos, P., and Kottommannil, J. (2003). “Practical
using other models of delay may yield a more universally appli- procedure for calibrating microscopic traffic simulation models.” Trans-
cable optimization model. Third, only the morning commute peak portation Research Record, 1852, 130–139.
period was modeled. As the A1 Highway is the only corridor along Jang, K., Chung, K., and Yeo, H. (2014). “A dynamic pricing strategy
the coast, peaks are common on the weekends when city dwellers for high occupancy toll lanes.” Transp. Res. Part A Policy Pract.,
visit homes in the mountains or in the summer when beach goers 67, 69–80.

© ASCE 04015029-10 J. Transp. Eng.

J. Transp. Eng.
Lave, C. A. (1969). “A behavioral approach to modal split forecasting.” Venglar, S. P., Fenno, D. W., Goel, S., and Schrader, P. A. (2002). “Man-
Transp. Res., 3(4), 463–480. aged lanes—Traffic modeling.” Rep. No. FHWA/TX-02/4160-4, Texas
Lindsey, R. (2012). “Road pricing and investment.” Econ. Transp., 1(1), Transportation Institute, College Station, TX.
49–63. Verhoef, E., Nijkamp, P., and Rietveld, P. (1996). “Second-best congestion
Nie, Y., and Liu, Y. (2010). “Existence of self-financing and Pareto-improv- pricing: The case of an untolled alternative.” J. Urban Econ., 40(3),
ing congestion pricing: Impact of value of time distribution.” Transp. 279–302.
Res. Part A, 44(1), 39–51. Wang, J. Y., Lindsey, R., and Yang, H. (2011). “Nonlinear pricing on pri-
Parry, I. W. (2009). “Pricing urban congestion.” Ann. Rev. Resour. Econ., vate roads with congestion and toll collection costs.” Transp. Res. Part
1(1), 461–484. B Method., 45(1), 9–40.
World Bank. (2013). “International comparison program database: World
Ross, P. (1991). “Some properties of macroscopic traffic models.” Trans-
development indicators.” 〈http://data.worldbank.org/indicator〉 (Jul. 20,
portation Research Record, 1194, 129–134.
2015).
Shires, J., and de Jong, G. (2009). “An international meta-analysis of values
Yang, H., and Huang, H.-J. (2004). “The multi-class, multi-criteria traffic
of travel time savings.” Eval. Program Plann., 32(4), 315–325. network equilibrium and systems optimum problem.” Transp. Res. Part
Small, K. A. (2012). “Valuation of travel time.” Econ. Transp., 1(1), 2–14. B, 38(1), 1–15.
Downloaded from ascelibrary.org by Cambridge University on 08/16/15. Copyright ASCE. For personal use only; all rights reserved.

Small, K. A., and Verhoef, E. T. (2007). The economics of urban transpor- Zhang, L., and Colyar, J., Pisano, P., and Holm, P. (2004). “Identifying and
tation, Taylor & Francis, London. assessing key weather related parameters and their impacts on traffic
Transportation Research Board. (2013). “Improving our understanding of operations using simulation.” Proc., Transportation Research Board
how highway congestion and pricing affect travel demand.” Strategic Annual Meeting 2005, Washington, DC.
Highway Research Program (SHRP 2) Rep. S2-C04-RW-1, 〈http:// Zhao, Z., An, S., and Wang, J. (2010). “Development and inspiration
www.trb.org/Main/Blurbs/168141.aspx〉 (Jul. 20, 2015). of road congestion pricing revenue redistribution theory research.” J.
TSIS-CORSIM [Computer software]. Gainesville, FL, McTrans Center. Transp. Syst. Eng. Inf. Technol., 10(4), 93–100.

© ASCE 04015029-11 J. Transp. Eng.

J. Transp. Eng.

Вам также может понравиться