Models for Evaluating Airline Overbooking 301
Models for Evaluating Airline
Overbooking
Michael P. Schubmehl
Wesley M. Turner
Daniel M. Boylan
Harvey Mudd College
Claremont, CA
Advisor: Michael E. Moody
Introduction
We develop two models to evaluate overbooking policies.
The first model predictstheeffectiveness of a proposedoverbookingscheme,
using the concept of expected marginal seat revenue (EMSR). This model solves
the discount seat allocation problem in the presence of overbooking factors for
each fare class and evaluates an overbooking policy stochastically.
The second model takes in historical flight data and reconstructs what the
optimal seat allocation should have been. The percentage of overbooking rev-
enue obtained in practice serves as a measure of the policy’s value.
Finally, we examine the overbooking problem in light of the recent drastic
changes to airline industry and conclude that increased overbooking would
bring short-term profits to most carriers. However, the long-term ill effects
that have traditionally caused airlines to shun such a policy would be even
more pronounced in a post-tragedy climate.
Literature Review
There are two major ways that airlines try to maximize revenues: over-
booking (selling more seats than available on a given flight) and seat allocation
(price discrimination). These measures are believed to save major airlines as
much as half a billion dollars each year, in an industry with a typical yearly
profit of about $1 billion dollars [Belobaba 1989].
The UMAP Journal 23 (3) (2002) 301–316.
c
Copyright 2002 by COMAP, Inc. All rights reserved.
Permission to make digital or hard copies of part or all of this work for personal or classroom use
is granted without fee provided that copies are not made or distributed for profit or commercial
advantage and that copies bear this notice. Abstracting with credit is permitted, but copyrights
for components of this work owned by others than COMAP must be honored. To copy otherwise,
to republish, to post on servers, or to redistribute to lists requires prior permission from COMAP.
302 The UMAP Journal 23.3 (2002)
Beckman [1958] models booking and no-shows in an effort to find an optimal
overbooking strategy. He ignores advance cancellations, assuming that all can-
cellations are no-shows [Rothstein 1985]. A method that is easy to implement
but sophisticated enough to allow for cancellations and group reservations was
developed by Taylor [1962]. Versions of this model were implemented at Iberia
Airlines [Shlifer and Vardi 1975], British Overseas Airways Corporation, and
El Al Airlines [Rothstein 1985].
None of these approaches considers multiple fare classes. Littlewood [1972]
offers a simple two-fare allocation rule: A discount fare should be sold only if
the discount fare is greater than or equal to the expected marginal return from
selling the seat at full-fare. This idea was generalized by Belobaba [1989] to in-
clude any number of fare classes and allow the integration of overbooking. We
use expected marginal seat revenue in predicting about overbooking schemes.
There is a multitude of work on the subject [McGill 1999]—according to
Weatherford and Bodily [1992], there are more than 124,416 classes of models
for variations of the yield management problem, though research has settled
into just a few of these. Several authors tried to better Belobaba’s [1987] heuristic
in the presence of three or more fare classes (for which it is demonstrably sub-
optimal) [Weatherford and Bodily 1992]; generally, these adaptive methods for
obtaining optimal booking limits for single-leg flights are done by dynamic
programming [Mcgill 1999].
After deregulation in 1978, airlines were no longer required to maintain a
direct-route system to major cities. Many shifted to a hub-and-spoke system,
and network effects began to grow more important. To maximize revenue, an
airline may want to consider a passenger’s full itinerary before accepting or
denying their ticket request for a particular leg. For example, an airline might
prefer to book a discount fare rather than one at full price if the passenger is
continuing on to another destination (and thus paying an additional fare).
The first implementations of the origin-destination control problem consid-
ered segments of flights. The advantage to this was that a segment could be
blacked out to a particular fare class, lowering the overall complexity of a book-
ing scheme. Another method, virtual nesting, combines fare classes and flight
schedules into distinct buckets [Mcgill 1999]. Inventory control on these buck-
ets would then give revenue-increasing results. Finally, the bid-price method
deterministically assigns a value to different seats on a flight leg. The legs in
an itinerary are then summed to establish a bid-price for that itinerary; a ticket
request is accepted only if the fare exceeds the bid-price [Mcgill 1999]
The most realistic yield management problem takes into account five price
classes. The ticket demands for different fare classes are randomized and corre-
lated with one other to allow for sell-ups and the recapture of rejected customers
on later flights. Passengers can no-show or cancel at any time. Group reser-
vations are treated separately from individuals—their cancellation probability
distribution is likely different. Currently, most work assumes that passengers
who pay full fare would not first check for availability of a lower-class ticket;
a more realistic model would allow buyers of a higher-class ticket to be di-
Models for Evaluating Airline Overbooking 303
verted by a lower fare. A full accounting of network effects would consider
the relative value of what Weatherford and Bodily [1992] terms displacement
denying a discount passenger’s ticket request to fly a multileg itinerary in favor
of leaving one of the legs open to a full-fare passenger.
Unfortunately, while the algorithms for allocating seats and setting over-
booking levels are highly developed, there has been little work done on the
problem of evaluating how effective these measures actually are. Our solution
applies industry-standard methods to find optimal booking levels, then exam-
ines the actual booking requests for a given flight to determine how close to an
optimal revenue level the scheme actually comes.
Factors Affecting Overbooking Policy
General Concerns
The reason that overbooking is so important is because of multiple fare
classes. With only one fare class, it would be easier for airlines to penalize
customers for no-shows. However, while most airlines offer nonrefundable
discount tickets, they prefer not to penalize those who pay full fare, like business
travelers, because these passengers account for most of the profits.
The overbooking level of a plane is dictated by the likelihood of cancellations
and of no-shows. An overbooking model compares the revenue generated by
accepting additional reservations with the costs associated with the risk of
overselling and decides whether additional sales are advisable. In addition,
the “recapture” possibility can be considered, which is the probability that a
passenger denied a ticket will simply buy a ticket for one of the airline’s other
flights. Since a passenger is more valuable to the airline buying a ticket on a
flight that has empty seats to fill than on one that is already overbooked, a high
recapture probability reduces the optimal overbooking level [Smith et al. 1992].
No major airline overbooks at profit-maximizing levels, because it could
not realistically handle the problems associated with all the overloaded flights.
This gives the overbooking optimization problem some important constraints.
The total flight revenue is to be maximized, subject to the conditions that only
a certain portion of flights have even one passenger denied boarding (one
oversale), and that a bound is placed on the expected total number of oversales.
Dealing with even one oversale is a hassle for the airline’s staff, and they are
not equipped to handle such problems on a large scale. Additionally, some
research indicates that increased overbooking levels would most likely trigger
an “overbooking war [Suzuki 2002], which would increase short-term profits
but would probably engender enough consumer resentment that the industry
as a whole would lose business.
While the overbooking problem sets a limit for sales on a flight as a whole,
proper seat allocation sets an optimal point at which to stop selling tickets
for individual fare levels. For example, a perfectly overbooked plane, loaded
304 The UMAP Journal 23.3 (2002)
exactly to capacity, could be flying at far below its optimal revenue level if
too many discount tickets were sold. The more expensive tickets are not for
first-class seats and involve no additional luxuries above the discount tickets,
apart from more lenient cancellation policies and the ability to buy the tickets
a shorter time before the flight’s departure.
September 11, 2001
Since the September 11 terrorist attacks, there have been significant changes
in the airline business. In addition to the forced cancellation of many flights in
the immediate aftermath of the attacks and the extreme levels of cancellations
and no-shows by passengers after that, passenger traffic has dropped sharply
in general. The huge downturn in passenger levels has led to large reductions
in service by most carriers.
In terms of the booking problem, there are fewer flights for passengers
to spill over onto, which could increase the loss by an airline if it bumps a
passenger from a flight. On the other hand, since passenger levels have fallen
so far, it is less likely that an airline will overfill any given flight. The heightened
security at airports will likely increase the passenger no-show rate somewhat,
as passengers get delayed at security checkpoints. At the very least, it should
almost completely remove the problem of “go-shows,” passengers who show
up for a flight but are not in the airline’s records.
On the whole, optimal booking strategies have become even more vital as
airlines have already lost billions of dollars, and some teeter on the brink of
failure. Some overbooking tactics previously dismissed as too harmful in the
long run might be worthwhile for companies in trouble. For example, an airline
near failure might increase the overbooking rate to the level that maximizes
revenue, without regard to the inconvenience and possible future resentment
of its customers.
Constructing the Model
Objectives
A scheme for evaluating overbooking policies needs to answer two ques-
tions: how well should a new overbooking method perform, and how well is a
current overbooking scheme already working? The first is best addressed by a
simple model of an airline’s booking procedures; given some setup for allocat-
ing seats to fare classes, candidate overbooking schemes can be laid on top and
tested by simulation. This approach has the advantage that it provides insight
into why an overbooking scheme is or is not effective and helps to illuminate
the characteristics of an optimal overbooking approach.
The second question is, in some respects, a simpler one to answer. Given the
actual (over)booking limits that were imposed on each fare class, and all avail-
Models for Evaluating Airline Overbooking 305
able information on the actual requests for reservations, how much revenue
might have been gained from overbooking, compared to how much actually
was? This provides a very tangible measure of overbooking performance but
very little insight into the reasons for results.
The enormous number of factors affecting the design and evaluation of
an overbooking policy force us to make simplifying assumptions to construct
models that meet both of these goals.
Assumptions
Fleet-wide revenues can be near-optimized one leg at a time.
Maximizing revenue involves complicated interactions between flights. For
instance, a passenger purchasing a cheap ticket on a flight into a major hub
might actually be worth more to the airline than a business-class passenger,
on account of connecting flights. We assume that such effects can be compen-
sated for by placing passengers into fare classes based on revenue potential
rather than on the fare for any given leg. This assumption effectively reduces
the network problem to a single-leg optimization problem.
Airlines set fares optimally.
Revenue maximization depends strongly on the prices of various classes
of tickets. To avoid getting into the economics of price competition and
supply/demand, we assume that airlines set prices optimally. This reduces
revenue maximization to setting optimal fare-class (over)booking limits.
Historical demand data are available and applicable.
The model needs to estimate future demand for tickets on any given flight.
We assume that historical data are available on the number of tickets sold
any given number of days t before a flight’s departure. In some respects, this
assumption is unrealistic because of the problem of data censorship—that is,
the failure of airlines to record requests beyond the booking limit for a fare
class [Belobaba 1989]. On the other hand, statistical methods can be used
to reconstruct this information [Boeing Commercial Airline Company 1982,
7–16; Swan 1990].
Low-fare passengers tend to book before high-fare ones.
Discount tickets are often sold under advance purchase restrictions, for the
precise reason that it enables price discrimination. Because of restrictions
like these, and because travelers who plan ahead search for cheap tickets,
low-fare passengers tend to book before high-fare ones.
Predicting Overbooking Effectiveness
Disentangling the effects of overbooking, seat allocation, pricing schemes,
and external factors on revenues of an airline is extremely complicated. To
306 The UMAP Journal 23.3 (2002)
isolate the effects of overbooking as much as possible, we want a simple, well-
understood seat allocation model that provides an easy way to incorporate
various overbooking schemes. In light of this objective, we pass up several
methods for finding optimal booking limits on single-leg flights detailed in, for
example, Curry [1990] and Brumelle [1993], in favor of the simpler expected
marginal seat revenue (EMSR) method [Belobaba 1989].
EMSR was developed as an extension of the well-known rule of thumb,
popularized by Littlewood [1972], that revenues are maximized in a two-fare
system by capping sales of the lower-class ticket when the revenue from selling
an additional lower-class ticket is balanced by the expected revenue from selling
the same seat as an upper-class ticket. In the EMSR formulation, any number of
fare classes are permitted and the goal is “to determine how many seats not to
sell in the lowest fare classes and to retain for possible sale in higher fare classes
closer to departure day” [Belobaba 1989].
The only information required to calculate booking levels in the EMSR
model is a probability density function for the number of requests that will
arrive before the flight departs, in each fare class and as a function of time.
For simplicity, this distribution can be assumed to be normal, with a mean and
standard deviation that change as a function of the time remaining. Thus, the
only information an airline would need is a historical average and standard
deviation of demand in each class as a function of time. Ideally, the informa-
tion would reflect previous instances of the particular flight in question. Let
the mean and standard deviations in question be denoted by μ
i
(t) and σ
i
(t) for
each fare class i =1, 2,... ,k. Then the probability that demand is greater than
some specified level S
i
is given by
¯
P
i
(S
i
,t)
1
2πσ
i
(t)
S
i
e
(rμ
i
(t))
2
i
(t)
2
dr.
This spill probability is the likelihood that the S
i
th ticket would be sold if offered
in the ith category. If we further allow f
i
(t) to denote the expected revenue
resulting from a sale to class i at a time t days prior to departure, we can define
EMSR
i
(S
i
,t)=f
i
(t) ·
¯
P
i
(S
i
,t),
or simply the revenue for a ticket in class i times the probability that the S
i
th
seat will be sold. The problem, however, is to find the number of tickets S
i
j
that
should be protected from the lower class j for sale to the upper class i (ignoring
other classes for the moment). The optimal value for S
i
j
satisfies
EMSR
i
(S
i
j
,t)=f
j
(t), (1)
so that the expected marginal revenue from holding the S
i
j
th seat for class i is
exactly equal to (in practice, slightly greater than) the revenue from selling it
immediately to someone in the lower class j. The booking limits that should
be enforced can be derived easily from the optimal S
i
j
values by letting the
Models for Evaluating Airline Overbooking 307
booking limit B
j
for class j be
B
j
(t)=C S
j+1
j
i<j
b
i
(t), (2)
that is, the capacity C of the plane, less the protection level of the class above j
from class j and less the total number of seats already reserved. Sample EMSR
curves, with booking limits calculated in this fashion, are shown in Figure 1.
EMSR ($)
10 20 30 40 50 60
0
50
100
150
200
250
b1b2
Seat Number
Figure 1. Expected marginal seat revenue (EMSR) curves for three class levels, with the high-
est-revenue class at the top. Each curve represents the revenue expected from protecting a par-
ticular seat to sell to that class. Also shown are the resulting booking limits for each of the lower
classes—that is, the levels at which sales to the lower class should stop to save seats for higher
fares.
This formulation does not account for overbooking; if we allow each fare
class i to be overbooked by some factor OV
i
, the optimality condition (1) be-
comes
EMSR
i
(S
i
j
,t)=f
j
(t) ·
OV
i
OV
j
. (3)
This can be understood in terms of an adjustment to f
i
and f
j
; the overbooking
factors are essentially cancellation probabilities, so we use each OV
i
to deflate
the expected revenue from fare class i. Then
¯
P
i
(S
i
j
,t) ·
f
i
(t)
OV
i
= f
j
(t) ·
f
j
(t)
OV
j
,
308 The UMAP Journal 23.3 (2002)
which is equivalent to (3). Note that the use of a single overbooking factor for
the entire cabin (that is, OV
i
= OV ) causes the OV
i
and OV
j
in (3) to cancel.
Nonetheless, the boarding limits for each class are affected, because the capacity
of the plane C must be adjusted to account for the extra reservations, so now
C
= C · OV
and the booking limits in (2) are adjusted upward by replacing C with C
.
The EMSR formalism gives us the power to evaluate an overbooking scheme
theoretically by plugging its recommendations into a well-understood stable
model and evaluating them. Given the EMSR boarding limits, which can be
updated dynamically as booking progresses, and the prescribed overbooking
factors, a simulated string of requests can be handled. Since the EMSR model
involves only periodic updates to establish limits that are fixed over the course
of a day or so, a set of n requests can be handled with two lookups each (booking
limit and current booking level), one subtraction, and one comparison; so all
n requests can be processed on O(n) time. An EMSR-based approach would
thus be practical in a real-world real-time airline reservations system, which
often handles as many as 5,000 requests per second. Indeed, systems derived
from EMSR have been adopted by many airlines [Mcgill 1999].
Evaluating Past Overbookings
The problem of evaluating an overbooking scheme that has already been
implemented is somewhat less well studied than the problem of theoretically
evaluating an overbooking policy. One simple approach, developed by Ameri-
can Airlines in 1992, measures the optimality of overbooking and seat allocation
separately [Smith et al. 1992]. Their overbooking evaluation process assumes
optimal seat allocation and, conversely, their seat allocation evaluation scheme
assumes optimal overbooking. Under this assumption, an overbooking scheme
is evaluated by estimating the revenue under optimal overbooking in two ways:
If a flight is fully loaded and no passenger is denied boarding, the flight
is considered to be optimally overbooked and to have achieved maximum
revenue.
If n passengers are denied boarding, the money lost due to bumping these
passengers is added back in and the n lowest fares paid by passengers for
the flight are subtracted from revenue.
On the other hand, if there are n empty seats on the plane, the n highest-fare
tickets that were requested but not sold are added to create the maximum
revenue figure.
Their seat-allocation model estimates the demand for each flight by calcu-
lating a theoretical demand for each fare class and then setting the minimum
flight revenue (by filling the seats lowest-class first) and the maximum flight
Models for Evaluating Airline Overbooking 309
revenue (by filling the seats highest-class first). To estimate demand, we use
the information on the flight’s sales up to the point where each class closed. By
assuming that demand is increasing for each class, we can project the number
of requests that would have occurred had the booking limits been disregarded.
Given these projected additional requests and the actual requests received
before closing, it is straightforward to compute the best- and worst-case over-
booking scenarios. The worst-case revenue R
is determined by using no
booking limits and taking reservations as they come, and the best-case rev-
enue R
+
is determined by accommodating high-fare passengers first, giving
the leftovers to the lower classes. The difference between these two figures is
the revenue to be gained by the use of booking limits. Thus, the performance
of a booking scheme that generates revenue R is
p =
R R
R
+
R
· 100%, (4)
representing the percentage of the possible booking revenue actually achieved.
We select this method for evaluating booking schemes after the fact.
Analysis of the Models
Tests and Simulations
The EMSR method requires information on demand as a function of time.
Although readily available to an airline, it is not widely published in a detailed
form. Li [2001] provides enough data to construct a rough piecewise-linear
picture of demand remaining as a function of time, shown in Figure 2.
This information can be inputted into the EMSR model to produce optimal
booking limits that evolve in time. A typical situation near the beginning of
ticket sales was shown in Figure 1, while the evolution of the limits themselves
is plotted in Figure 3.
The demand information in Figure 2 can also be used to simulate requests
for reservations. By taking the difference between the demand remaining at
day t and at day (t 1) before departure, the expected demand on day t can be
determined. The actual number of requests generated on that day is then given
by a Poisson random variable with parameter λ equal to the expected number
of sales [Rothstein 1971]. The requests generated in this manner can be passed
to a request-handling simulation that looks at the most current booking limits
and then accepts or denies ticket requests based on the number of reservations
already confirmed and the reservations limit. An example of this booking
process is illustrated in Figure 4.
The results of the booking process provide an ideal testbed for the revenue
opportunity model employed to evaluate overbooking performance. The sim-
ulation conducted for Figure 4 had demand values of {11, 41, 57}, for classes 1,
310 The UMAP Journal 23.3 (2002)
Expected Remaining Demand as Flight Time Approaches
Fraction of Demand Remaining
80
60 40 20 0
0.2
0.4
0.6
0.8
1
Days Until Departure
Figure 2. Demand remaining as a function of time for each of three fare classes, with the highest
fare class on top. The curves represent the fraction of tickets that have yet to be purchased. Note
that, for example, demand for high fare tickets kicks in much later than low-fare demand. ( Data
interpolated from Li [2001].)
2, and 3 respectively, before ticket sales were capped. A linear forward projec-
tion of these sales rates indicates that they would have reached {18, 49, 69}had
the classes remained open. Given fare classes {$250, $150, $100}, the minimum
revenue would be
R
= $100(69) + $150(40) + $250(0) = $12 , 900
and the maximum revenue would be
R
+
= $250(18) + $150(49) + $100(42) = $16 , 050.
The actual revenue according to the EMSR formalism was
R = $100(57) + $150(41) + $250(11) = $14, 600,
so the efficiency is
p =
R R
R
+
R
· 100% =
$14, 600 $12, 900
$16, 050 $12, 900
· 100% = 54%,
without the use of a complicated overbooking scheme. This is not close to the
efficiencies reported in Smith et al. [1992], which cluster around 92%. This rela-
tive inefficiency is to be expected, however, from a simplified booking scheme
given incomplete booking request data.
Models for Evaluating Airline Overbooking 311
Evolution of Booking Limits by the EMSR Method
Booking Limit
80
60 40 20 0
20
40
60
80
100
Days Until Departure
Figure 3. Booking limits for each class are dynamically adjusted to account for tickets already sold.
For illustrative purposes, the number of tickets already sold is replaced here with the number
of tickets that should have been sold according to expectations. In this case, the booking limits
estimated at the beginning of the process are fairly accurate and require relatively little updating.
Total Bookings by Fare Class: First Sale to Flight Time
Bookings
80
60 40 20 0
10
20
30
40
50
Days Until Departure
Figure 4. The EMSR-based booking limits are used to decide whether to accept or reject a sequence
of ticket requests. These requests follow a Poisson distribution where the parameter λ varies with
time to match the expected demand. Each fare class reaches its booking limit, as desired, so the
flight is exactly full. Incorporating overbooking factors shifts the limits up accordingly.
312 The UMAP Journal 23.3 (2002)
Strengths and Weaknesses
Strengths
Applies widely accepted, industry-standard techniques.
Although more advanced (and optimal) algorithms are available and are
used, EMSR and its descendants are still widely used in industry and can
come close to optimality. The EMSR scheme, tested as-is on a real airline,
caused revenue gains as much as 15% [Belobaba 1989].
Our method for determining the optimality of a scheme after the fact is also
based on tried and true methods developed by American Airlines [Smith et
al. 1992].
Simplicity
Since it does not take into account as many factors as other booking models,
EMSR is easier to deal with computationally. While a simple model may not
be able to model a major airline with complete accuracy, an optimal pricing
scheme can be made using only three fare classes [Li 2001].
Weaknesses
Neglects network effects
We treat the problem of optimizing each flight as if it were an independent
problem although it is not.
Ignores sell-ups
In considering the discount seat allocation problem, we treat the demands
for the fare classes as constants, independent of one other. This is not the
case, because of the possibility of sell-ups. If the number of tickets sold in a
lower fare class is restricted, then there is some probability that a customer
requesting a ticket in that class will buy a ticket at a more expensive fare.
This means it is possible to convert low-fare demand into high-fare demand,
which would suggest protecting a higher number of seats for high fares than
calculated by the model that we use. Sell-ups would be straightforward to
incorporate into EMSR, but doing so would require additional information
[Belobaba 1989].
Discounts possibility of recapture
Similar to sell-ups, the recapture probability is the probability that a passen-
ger unable to buy a ticket at a certain price on a given flight will buy a ticket
on a different flight. Depending on the recapture probability for each fare
class, more or fewer seats might be allocated to discount fares.
Models for Evaluating Airline Overbooking 313
Recommendations on Bumping Policy
In 1999, an average of only 0.88 passengers per 10,000 boardings were in-
voluntarily bumped. Airlines are not required to keep records of the number
of voluntary bumps, so it is impossible to determine a general bump rate.
Before bumping passengers involuntarily, the airline is required to ask for
volunteers. Because there are no regulations on compensation for voluntary
bumps, this is often a cheaper and more attractive method for airlines anyway.
If too few people volunteer, the airline must pay those denied boarding 200%
of the sum of the values of the passengers’ remaining flight coupons, with a
maximum of $400. This maximum is decreased to $200 if the airline arranges
for a flight that will arrive less than 2 hours after the original flight. The air-
line may also substitute the offer of free or reduced fare transportation in the
future, provided that the value of the offer is greater than the cash payment
otherwise required. Alternatively, the airline may simply arrange alternative
transportation if it is scheduled to arrive less than an hour after the original
flight.
Auctions in which the airline offers progressively higher compensation for
passengers who give up their seats are both the cheapest and the most com-
mon practice. As long as the airline does not engage in so much overbooking
that it cannot find suitable reroutes for passengers bumped from their original
itineraries, no alternatives to this policy need to be considered.
Conclusions
The two models presented in this paper work together to evaluate over-
booking schemes by simulating their effects in advance and by quantifying
their effects after implementation.
The expected marginal seat revenue (EMSR) model predicts overbooking
scheme effectiveness. It determines the correct levels of protection for each fare
class above the lowest—that is, how many seats should be reserved for possible
sale at later dates and higher fares. Overbooking factors can be specified sepa-
rately for each fare class, so the model effectively takes in overbooking factors
and produces booking limits that can be used to handle ticket requests.
The revenue opportunity model attempts to estimate the maximum revenue
from a flight under perfect overbooking and discount allocation. This is accom-
plished by estimating the actual demand for seats, then calculating the revenue
that these seats would generate if sold to the highest-paying customers. Sim-
ple calculations produce the ideal overbooking cap and the optimal discount
allocation for the flight. Thus, this model effectively represents how the airline
would sell tickets if they had perfect advance knowledge of demand.
After the September terrorist attacks and their subsequent catastrophic ef-
fects on the airline industry, heightened airport security and fearful passengers
will increase no-show and cancellation rates, seeming to dictate increasing
314 The UMAP Journal 23.3 (2002)
overbooking levels to reclaim lost profits.
Airlines considering such action should be cautioned, however, that the
negative effects of increased overbooking could outweigh the benefits. With
reduced airline service, finding alternative transportation for displaced passen-
gers could be more difficult. The effect of denying boarding to more passen-
gers, along with the greater inconvenience of being bumped, could seriously
shake consumers’ already-diminished faith in the airline industry. With air-
lines already losing huge numbers of customers, it would be a mistake to risk
permanently losing them to alternatives like rail and auto travel by alienating
them with frequent overselling.
References
Beckman, J.M. 1958. Decision and team problems in airline reservations. Econo-
metrica 26: 134–145.
Belobaba, Peter. 1987. Airline yield management: An overview of seat inven-
tory control. Transportation Science 21: 63–73.
. 1989. Application of a probabilistic decision model to airline seat
inventory control. Operations Research 37: 183–197.
Biyalogorsky, Eyal, et al. 1999. Overselling with opportunistic cancellations.
Marketing Science 18 (4): 605–610.
Boeing Commercial Airline Company. 1982. Boeing Promotional Fare Man-
agement System: Analysis and Research Findings. Seattle, WA: Boeing,
Seattle.
Bowen, B.D., and D.E. Headley. 2000. Airline Quality Rating 2000. http:
//www.uomaha.edu/~unoai/research/aqr00/ .
Brumelle, S.L., and J.I. McGill. 1993. Airline seat allocation with multiple
nested fare classes. Operations Research 41: 127–137.
Chatwin, Richard E. 2000. Optional dynamic pricing of perishable products
with stochastic demand and a finite set of prices. European Journal of Oper-
ations Research 125: 149–174.
Curry, R.E. 1990. Optimal airline seat allocation with fare classes nested by
origins and destinations. Transportation Science 24: 193–204).
Davis, Paul. Airline ties profitability yield to management. 1994. Siam News
27 (5); http://www.siam.org/siamnews/mtc/mtc694.htm .
Li, M.Z.F. 2001. Pricing non-storable perishable goods by using a purchase
restriction with an application to airline fare pricing. European Journal of
Operations Research 134: 631–647.
Models for Evaluating Airline Overbooking 315
Littlewood, K. 1972. Forecasting and control of passenger bookings. AGIFORS
Symposium Proceedings: 12.
McGill, J.I., and G.Z. and Van Ryzin. 1999. Revenue management: Research
overview and prospects. Transportation Science 33 (2): 233–256.
Rothstein, M. 1971. An airline overbooking model. Transportation Science 5:
180–192.
Rothstein, Marvin 1985. OR and the airline overbooking problem. Operations
Research 33 (2): 237–247.
Ross, Robert. 1998. Embry-Riddle experts reveal airfare secrets. http://comm.
db.erau.edu/leader/fall98/priced.html .
Shlifer, R., and Y. Vardi. 1975. An Airline Overbooking Policy. Transportation
Science 9: 101-114.
Swan. 1990. Revenue management forecasting biases. Working paper. Seattle,
WA: Boeing Commercial Aircraft.
Smith, Barry C., et al. 1992. Yield management at American Airlines. Interfaces
22 (1): 8–31.
Suzuki, Yoshinori. 2002. An empirical analysis of the optimal overbooking
policies for U.S. major airlines. Transportation Research E38: 135–149.
Taylor, C.J. 1962. The determination of passenger booking levels. AGIFORS
Symposium Proceedings, vol. 2. Fregene, Italy.
Weatherford, L.R., and S.E. Bodily. 1992. A taxonomy and research overview of
perishable-asset revenue management: Yield management, overbooking,
and pricing. Operations Research 40: 831–844.
Presentation by Richard Neal (MAA Student Activities Committee Chair) of the MAA award to
Daniel Boylan and Wesley Turner of the Harvey Mudd College team (Michael Schubmehl could
not come), after their presentation at the MAA Mathfest in Burlington, VT, in August. On the right
is Ben Fusaro, Founding Director of the MCM. Photo by Ruth Favro, Lawrence Tech University.
316 The UMAP Journal 23.3 (2002)
Letter to the CEO of a Major Airline
Airline overbooking is just one facet of a revenue management problem that
has been studied extensively in operations research literature. Airlines have
been practicing overbooking since the 1940’s, but early models of overbooking
considered only the most rudimentary cases. Most importantly, they did not
take into account the revenue maximizing potential of price discrimination—
charging different fares for identical seats. In order to maximize yield, it is
particularly critical to price discriminate between business and leisure travelers.
That is, when filling the plane, book as many full fare passengers and as few
discount fare passengers as possible.
The implementation of a method of yield management can have dramatic
effects on an airline’s revenue. American Airlines managed its seat inventory
to a $1.4 billion increase in revenue from 1989 to 1992—about 50% more than
its net profit for the same period. Controlling the mix of fare products can
translate into revenue increases of $200 million to $500 million for carriers with
total revenues of $1 billion to $5 billion.
Though several decision models of airline booking have been developed
over the years, comparing one scheme to another remains a difficult task. We
have taken a two-pronged approach to this problem, both simulating and mea-
suring a booking scheme’s profitability.
In orderto simulate a booking scheme’s effect, we used the expected marginal
seat revenue (EMSR) model proposed by Belobaba [1989] to generate near-
optimal decisions on whether to accept or deny a ticket request in a given fare
class. The EMSR model accepts as input overbooking levels for each of the fare
classes that compose a flight, so different policies can be plugged in for testing.
Our approach to measuring a current scheme’s profitability is similar to one
used at American Airlines [Smith et al. 1992]. We compare the actual revenue
generated by a flight with an ideal level calculated with the benefit of hindsight,
as well as with a baseline level that would have been generated had no yield
management been used. By calculating the percentage of this spread earned by
a flight employing a particular scheme, we are able to gauge the effectiveness
of different booking schemes.
It is our hope that these models will prove useful in evaluating your airline’s
overbooking policies. Simulations should provide insight into the properties of
an effective scheme, and measurements after the fact will help to provide perfor-
mance benchmarks. Finally, while it may be tempting to increase overbooking
levels in order to compensate for lost revenues in the post-tragedy climate, our
results indicate this will probably hurt long-term profits more than they will
help.
Cordially,
Michael P. Schubmehl, Wesley M. Turner, and Daniel M. Boylan