ChapterPDF Available

Operating Theatre Planning and Scheduling

Authors:

Abstract and Figures

In this chapter we present a number of approaches to operating theatre planning and scheduling. We organize these approaches hierarchically which serves to illustrate the breadth of problems confronted by researchers. At each hierarchicalplanning level we describe common problems, solution approaches and results from studies at partner hospitals.
Content may be subject to copyright.
Chapter 5 Operating Theatre Planning and
Scheduling
Erwin W. Hans, Peter T. Vanberkel
Center for Healthcare Operations Improvement & Research, dep. Operational Methods for
Production & Logistics, University of Twente, Enschede
Abstract In this chapter we present a number of approaches to operating theatre
planning and scheduling. We organize these approaches hierarchically which
serves to illustrate the breadth of problems confronted by researchers. At each hi-
erarchical planning level we describe common problems, solution approaches and
results from studies at partner hospitals.
Acknowledgments The hospitals Erasmus MC and Netherlands Cancer Institute, and all co-
authors involved in the various research projects described here: Boucherie RJ, Harten WH van,
Houdenhoven M van, Hulshof PJH, Hurink JL, Kazemier G, Lans M van der, Lent WAM van,
Oostrum JM van, Wullink G.
5.1 Introduction
Within the OR/OM healthcare literature, operating theatre planning and schedul-
ing is one of the most popular topics. This is not surprising, as many patients in a
hospital undergo surgical intervention in their care pathway. For a hospital, the
operating theatre accounts for more than 40% of its revenues and a similar large
part of its costs (HFMA 2005). An efficient operating theatre department thus sig-
nificantly contributes to an efficient healthcare delivery system as a whole.
An extensive overview and taxonomy of the operating theatre planning and
scheduling literature is given by Cardoen et al. (2010). They conclude that the ma-
jority of the research is directed at planning and scheduling of elective patients at
an operational level of control, and take a deterministic approach. Furthermore,
they observe that only half of the literature contributions consider up- or down-
stream hospital resources, and few papers report about implementation in practice.
This appears to be a common problem in OR/OM healthcare literature (Brailsford
et al. 2009). An up-to-date online bibliography of the operating theatre manage-
ment literature is maintained by Dexter (2011), and a structured literature review
of operations research in the management of operating theatres is given by Guerri-
ero and Guido (2011).
2
In this chapter we address operating theatre planning problems on three hierar-
chical managerial levels: strategic, tactical and offline operational planning, as in-
troduced in Section 5.2.
The remainder of this chapter addresses recent work in each of these three lev-
els of control. Section 5.2 outlines the planning and control functions on the
aforementioned hierarchical levels in an operating theatre department. Section 5.3
addresses the strategic problem of determining the target utilization of an operat-
ing theatre department. Section 5.4 addresses the strategic problem of determining
the number of surgical teams required during the night to deal with emergency
cases. Section 5.5 addresses the strategic decision whether to use emergency oper-
ating theatres. Section 5.6 addresses the tactical problem of determining a master
surgery schedule (a day-to-day allocation of operating theatres to surgical special-
ties) that levels the workload in subsequent departments (wards). Section 5.7 ad-
dresses the offline operational problem of scheduling elective surgeries with sto-
chastic durations, and sequencing them in order to reduce access time of
emergency surgeries. We will use a wide array of OR techniques, including dis-
crete event and Monte Carlo simulation, statistical modeling and meta-heuristics.
5.2 A hierarchy of resource planning and control in operating
theatres
Competitive manufacturing companies make planning and control decisions in a
hierarchical manner (Zijm 2000). For example, the long term decision of what
products to manufacture is at the top of the hierarchy and the real time decision of
whether to discard a specific part due to its quality is at the bottom of the hierar-
chy. In general the reliance of one decision on another defines the hierarchy.
Many planning and control frameworks classify decisions into the three hierar-
chical levels strategic, tactical, and operational, as suggested by Anthony (1965).
Similar hierarchical planning and control frameworks have been proposed for
healthcare (see Hans et al. 2011). Hans et al. refine the classical hierarchy by split-
ting the operational level into an offline and online operational level, where the
former is the in advance short term decision making, and the latter the monitoring
and control of the process in real time. In the remainder of this section we outline
the main operating theatre planning and control functions on these four hierar-
chical levels.
5.2.1 Strategic planning and control
To reach organizational goals, the strategic level addresses the dimensioning of
core OT resources, such as the number of OTs, the amount of personnel, instru-
ments (e.g. X-ray machines), etc. It also involves case mix planning, i.e. the selec-
3
tion of surgery types, and the determination of the desired patient type volumes
(Vissers et al. 2001). Agreements are made with surgical services / specialties
concerning their annual patient volumes and assigned OT time. The dimensioning
of subsequent departments’ resources (e.g. ward beds) is also done (Vanberkel and
Blake 2007). Strategic planning is typically based on historical data and/or fore-
casts. The planning horizon is typically long term, e.g. a year or more.
5.2.2 Tactical planning and control
The tactical level addresses resource usage over a medium term, typically with a
planning horizon of several weeks (Blake and Donald 2002, Wachtel and Dexter
2008). The actual aggregate patient demand (e.g. waiting lists, appointment re-
quests for surgery) is used as input. In this stage, the weekly OT time is divided
over specialties or surgeons, and patient types are assigned to days. For the divi-
sion of OT time, two approaches exist (Denton et al. 2010). When a closed block
planning approach is used, each specialty will receive a number of OT blocks
(usually OT-days). In an (uncommon) open block planning approach, OT time is
assigned following the arrival of requests for OT time by surgeons.
On the tactical level, the surgery sequence is usually not of concern. In-
stead, on this level it is verified whether the planned elective surgeries cause re-
source conflicts for the OT, for subsequent departments (ICU, wards), or for re-
quired instruments with limited availability (e.g. X-ray machines). The design of a
Master Surgical Schedule is a tactical planning problem.
5.2.3 Operational planning and control (offline)
The offline operational level addresses scheduling of specific patients to resources
(and as a consequence, the sequencing of activities) and typically involves a plan-
ning horizon of a week. It encompasses the rostering of OT-personnel, and reserv-
ing resources for add-on surgeries (Dexter et al. 1999). In addition, it addresses the
sequencing of surgeries (Denton et al. 2007), to prevent critical resource conflicts,
e.g. regarding X-ray machines, instrument sets, surgeons, etc. When there are no
dedicated emergency OTs, the sequencing of the elective surgeries can also aid in
spreading the planned starting times of elective surgeries (which are “break-in
moments” for emergency surgeries) in order to reduce the emergency surgery
waiting time (Wullink et al. 2007).
4
5.2.4 Operational planning and control (online)
The online operational level addresses the monitoring and control of the day-to-
day activities in the OT. Obviously at this level of control, all uncertainty materi-
alizes and has to be dealt with. If necessary, surgeries are rescheduled, or even
cancelled (Dexter et al. 2004, McIntosh et al. 2006). This is usually done by a day
coordinator in the OT department. Emergency surgeries, which have to be dealt
with as soon as possible, are scheduled, and emergency OT teams may have to be
assembled and dispatched to the first available OT. If there are emergency OTs,
these emergency surgeries are dispatched to these OTs. If there are no such OTs,
they are scheduled within the elective surgical schedule.
In summary, strategic planning typically addresses capacity dimensioning de-
cisions, considering a long planning horizon of multiple years. Tactical planning
addresses the aggregate capacity allocation to patient types, on an intermediate
horizon of weeks or months. Offline operational planning addresses the in-
advance detailed capacity allocation to elective patients, with a short planning
horizon of days and up to a few weeks. Online operational planning addresses the
monitoring and control of the process during execution, and encompasses for ex-
ample reacting to unforeseen events.
5.3 Strategic: the problem with using target OT utilization levels
Utilization of operating theatres is high on the agenda of hospital managers and
researchers and is often used as a measure of efficiency, both introspectively as
well as in benchmarking against other OT departments. As a result, much effort is
spent trying to maximize OT utilization and sometimes, without understanding the
factors affecting it. Using straightforward statistical analysis we show how the tar-
get OT utilization of a hospital depends on the patient mix and the hospital’s will-
ingness to accept overtime. This work is described in detail in Houdenhoven et al.
(2007). Similarly, the erroneousness of target ward occupancies is studied by Har-
per and Shahani (2002) and discussed by Green in Hall (2006).
5.3.1 General model
There are various ways to compute the utilization rate. We define the OT utiliza-
tion as the expected total surgery duration (including changeover/cleaning time)
divided by the amount of time allotted (see Figure 5.1):

 (5.1)
5
Our approach can be easily extended to deal with more extensive definitions of
OT utilization.
Fig. 5.1 Timeline for surgical cases
The amount of allotted time is computed as follows:
5.2
where the slack time (reserved capacity) is determined in such a way that a certain
frequency of overtime is achieved. This is a managerial choice: slack time reduces
cancellations and/or costly overtime, but also reduces OT utilization. The frequen-
cy of overtime depends on the distribution of the total surgery duration and can be
computed according to:

5.3
Now more formally, let and denote the average and standard deviation of
elective surgical case durations of type s, and let denote the number of cases
completed in one block. A type s may correspond with the surgeries of for exam-
ple a surgical specialty or a specific surgeon. Likewise, , and denote the
same for emergency cases. All these parameters are based on historical data. It fol-
lows that the total expected duration of all elective cases in one OT block is
 and the standard deviation of the total duration of these cases equals
. Accordingly:
 (5.4)
.. (5.5)
The accepted risk (or frequency) of overtime is denoted by . Now we can com-
plete equation (5.2). The amount of allotted time required to achieve an overtime
frequency of can be computed as follows:
1
st
electivecase 2
nd
electivecase
expectedtotalsurgeryduration
allottedtime
slack
caseduration caseduration caseduration
8:00 15:30
3
rd
electivecase
6 5.6
where is a function yielding probability . The outcome of this function de-
pends on the distribution of the surgery duration. Using in this way allows
the approach to be independent of the surgery duration distribution, i.e. function
can be changed to reflect various distributions. Using equations (5.4) and
(5.6) we can complete formula (5.1) for the expected OT utilization as a function
of the frequency of overtime as follows:

 (5.7)
5.3.2 General results
We use formula (5.7) for the expected OT utilization to illustrate the relationship
between OT utilization, patient mix and overtime frequency. In a theoretical sce-
nario where there is no surgery duration variability (i.e., 0), the ex-
pected OT utilization is obviously 100%.
As a case study we consider Erasmus Medical Center in Rotterdam, the
Netherlands. OT management in this hospital accepts a 30% risk of overtime. For
simplicity, they assume that the total surgery duration follows a normal distribu-
tion ~,. Using straightforward sta-
tistical analysis we can show that 0.5 when the acceptable frequency of
overtime is 30%.This is shown as follows. Let X be the total surgery duration,
then:
30%⟺
0.7⟺

0.7⟺
0.7
where ~0,1. It follows that 0.5.
We use 2 years of historical data from the aforementioned hospital. We
consider three different surgical specialties (i.e. three different patient mixes) and
for each we show the trade-off between expected OT utilization and overtime fre-
quency. This is illustrated in Figure 5.2.
7
Fig. 5.2 Trade-off between overtime probability and expected utilization
The calculated expected OT utilization can also be regarded as a target utilization,
or benchmark. Figure 5.2 shows that a single OT utilization target will result in
different overtime frequencies for each specialty. For example, a target utilization
of 80% will result in an overtime frequency of approximately 12% for ophthal-
mology but an overtime frequency of approximately 35% for ENT. In general, a
low risk of overtime and a complex patient mix will result in a low utilization rate.
If the accepted risk of overtime is higher and the patient mix less complex, then a
higher utilization can be achieved. Given that overtime is expensive (and perhaps
limited by collective bargaining agreements), this example illustrates the inade-
quacy of a single target OT utilization as a performance metric. It also illustrates
the importance of taking case mix characteristics into account when comparing
utilization figures between different OT departments.
5.4 Strategic: on-call or in-house nurses for overnight coverage
for emergency cases?
Treating emergency patients is a common activity for most hospitals. Likewise,
the operating theatre must be available to provide emergency operations 24 hours
per day. The night shift (e.g., from 11:00 pm to 7:30 am) is typically the most ex-
pensive shift to staff due to collective labor agreements and the inconvenient
hours. Determining minimum cost staffing levels that provide adequate coverage
to meet emergency demand is a strategic problem. In this section we describe a
30%
40%
50%
60%
70%
80%
90%
100%
0,1%
10,0%
20,0%
30,0%
40,0%
50,0%
EarNoseThroat
Ophthalmology
Orthopedics
Overtime probability(%)
Expectedutilization(%)
8
case study to determine appropriate night shift staffing levels at Erasmus Medical
Center. The outcomes of the study were successfully implemented.
5.4.1 General problem formulation
Covering the night shift is usually accomplished by using in-hospital and on-call
nurses. The in-hospital nurses are stationed in the hospital while waiting for emer-
gency cases. The on-call nurses wait at their homes for emergency cases (typically
there is a requirement that they can be present in the hospital within e.g. 30
minutes of being requested to do so) and are typically cheaper than in-hospital
teams. In general, A single nurse can complete exactly one case at a time but can
complete any number of cases in series until the end of the shift. The decision re-
quired in this problem is to determine how many in-hospital and on-call nurses are
necessary to meet the demand for emergency cases. Timeliness is of the essence
here, as these emergency cases may be very urgent.
When the first emergency case presents for surgery during a night shift,
in-house nurses respond. Depending on the hospital policy and the total number of
in-house nurses, an on-call nurse may be called in. In other words, some hospitals
may wait until 1, 2, … or all in-house nurses are busy before calling in a nurse
from home, while other hospitals may wait until all in-house nurses are busy and
an emergency case is present. For each subsequent emergency case this process is
repeated. Note that nurses are available to complete multiple surgeries per night
and are available again after completing a surgery. Finally surgeries cannot be
preempted. In this subsection we assume the hospital’s policy for calling an on-
call nurse is fixed, although determining this policy is, in and of itself, and inter-
esting research question.
There are generally two types of emergency cases: those that need to be
started immediately and those that can be delayed before being started. The former
we refer to as emergent cases and the latter as urgent cases. The acceptable delay,
or safety interval, for starting an urgent case varies: “for example a facility may
consider it imperative for a patient with a ruptured abdominal aortic aneurysm to
be operated on within 30 min of arrival, while a patient with an amputated finger
should be operated on within 90 min of arrival, and a patient with a perforated
gastric ulcer should be operated on within 3 h of arrival” (Oostrum et al. 2008).
By incorporating the acceptable delays for urgent cases it is possible to
postpone urgent case demand to a later cheaper shift, and/or postpone the case un-
til busy in-house nurses are free. To examine these possibilities in detail, Oostrum
et al. (2008) use a discrete-event simulation and a case study at Erasmus Medical
Center Rotterdam (Erasmus MC). To illustrate the benefits of postponing surger-
ies, the authors compare results with surgery postponements with the approach of
Dexter and O’Neill (2001) where surgery postponements are not used. In the fol-
lowing subsection we provide an overview of the results.
9
5.4.2 General results
Current practice at Erasmus MC had a team composition of 9 in-house nurses and
2 on-call nurses. Using the approach of Dexter and O’Neill, a team of 8 in-house
nurses and 2 on-call nurses was determined to be appropriate. A number of other
team compositions were considered, ranging from a total of 11 nurses to a total of
6. Each team composition represented a what-if scenario in the simulation model.
The simulation model was used to determine the number of surgeries cases start-
ing later than required.
To compute the cost of each team composition, observe that – since the
number of working hours does not depend on the team composition – we only
need to look at the cost of idle staff. Under Dutch law, the costs for nurses who
wait during the night shift are 107.5% of the regular hourly daytime wage for in-
house nurses, and 106% for nurses on call. We thus compute the cost of waiting
nurses in each team composition as follows:
numberofin‐housenurses1.075
numberofon‐callnurses1.06hourlywage
Figure 5.3 displays the cost of waiting and percentage of surgeries starting late for
the considered team compositions, where we assumed for simplicity that a regular
hour’s wage is 1. It shows that current practice of 9 in-house and 2 on-call nurses
performs the best. However, the waiting cost can be decreased by approximately
18.5% by switching to a team composition of 5 in-house and 4 on-call nurses, at
the expense of a 2% increase of late starts.
Fig. 5.3 Cost of waiting and late surgery starts for various team compositions
6
7
8
9
10
11
12
10% 12% 14% 16% 18% 20% 22% 24%
Costofwaiting(hourlywage=1)
Percentageofsurgeriesstartinglate
CurrentPractice:9inhouse,2oncall
Dexter &O'Neill:8inhouse,2oncall
7inhouse,2oncall
5inhouse,4oncall
4inhouse,4oncall
4inhouse,3oncall
4inhouse,
2oncall
10
For policy making, managers can use results like these to see the relative perfor-
mance cost associated with each staff assignment. The decision autonomy remains
with the policy makers and they are left to determine if cheaper staffing levels jus-
tify the a decrease in performance.
For more extensive results, we refer to (Oostrum et al. 2008),where the
authors present the distribution of cases starting later than required, surgical spe-
cialty specific results, results for multiple nurse types, and an extensive sensitivity
analyses. The sensitivity analyses showed that the approach can be generalized for
use in other centers.
Oostrum et al. (2008) report that heavy involvement of clinical staff in
this project was essential for the following reasons. Staff assessed the safety inter-
vals for urgent patients to ensure changes did not negatively affect patient’s safety.
They validated the discrete event simulation model, and suggested various scenar-
ios for sensitivity analyses. The visualizations provided by the computer simula-
tion aided to convince them of the final conclusions. As a result, despite the nega-
tive impact on their salary, the staff accepted the adjustment of the team
composition to 5 in-house and 4 on-call nurses. For Erasmus MC this intervention
resulted in an annual cost saving of 275,000 Euro.
5.5 Strategic: Emergency operating theatres or not?
During regular working hours, most hospitals either perform emergency opera-
tions in dedicated emergency OTs, or in regular elective patient OTs. For the se-
cond option a certain amount of slack is scheduled in order to fit in emergency
cases without causing excessive cancellations of elective cases. The choice to use
Policy 1 (reserving capacity in dedicated emergency OTs) or Policy 2 (reserving
capacity in multiple regular emergency OTs) is the strategic decision addressed in
this section. The difference between these two policies is illustrated graphically in
Figure 5.4.
11
Fig. 5.4 Cost of waiting and late surgery starts for various team compositions
The flow of patients is summarized as follows: “Emergency patients arriving at a
hospital that has adopted the first policy, will be operated immediately if the dedi-
cated OT is empty and will have to queue otherwise, whereas patients arriving at a
hospital that has adopted the second policy can be operated once one of the ongo-
ing elective cases has ended. Other planned cases will then be postponed to allow
the emergency operation” (Wullink et al. 2007).
Policy 1 has the advantage that the first emergency case of the day can
begin without delay, but all following cases may be subject to delay. Furthermore
this policy means only the emergency OTs needed to be equipped to deal with
emergency cases. Finally, as a result of emergency surgeries elective surgeries
will experience no delay (Bhattacharyya et al. 2006, Ferrand et al. 2010) and elec-
tive OTs will experience no overtime (Wixted et al. 2008).
Policy 2 cannot guarantee any emergency case will begin without delay,
but since emergency cases can be completed in more OTs, an opening (i.e. a case
finishing) for the subsequent cases may a happen sooner than in policy 1. The
benefits from this policy essentially result from flexibility. To ensure this flexibil-
ity (and the corresponding benefits) multiple (or all) of the OTs must be equipped
to deal with emergency cases.
5.5.1 General problem description
The decision that is required is to determine how to reserve OT capacity for emer-
gency cases, i.e. according to policy 1 or policy 2. There are advantages and dis-
advantages of both policies introduced above. Due to the stochastic nature of
emergency cases (arrivals and surgery durations) choosing the best policy is not
Policy 1:
Emergency OTs Reserved
OT time for
emergency
surgery
Scheduled
elective
surgeries
Policy 2:
Elective OTs
OT1 OT2 OT3 OT4 OT5 OT6 OT7 OT8
OT1 OT2 OT3 OT4 OT5 OT6 OT7 OT8
12
immediately obvious. To compare the policies we suggest evaluating the follow-
ing metrics:
- emergency surgery waiting time: the total delay, or the delay past what is
allowed to receive emergency surgery.
- elective surgery waiting time: the difference between the planned and ac-
tual starting time of an elective surgery.
- OT overtime: the time used for surgical procedures after the regular block
time has ended.
- OT utilization: the ratio between the total used operating time for elective
procedures and the available regular time.
The following instance parameters are taken into account: elective surgery volume
and duration characteristics, emergency surgery arrival and duration characteris-
tics.
5.5.2 General results
We summarize a case study (presented in detail in Wullink et al. 2007) where dis-
crete event simulation was used to prospectively evaluate both policies. The case
study was used to support decision making at Erasmus MC. When applying Policy
2, the hospital decided that all of their 12 OTs would be equipped to handle emer-
gency cases. In policy 1, with emergency capacity allocated to 1 dedicated emer-
gency OT, the remaining free OT time is allocated exclusively to elective OTs. In
policy 2, with emergency time allocated to each elective OT, the reserved OT time
is distributed evenly over all elective OTs. Figure 5.5 and Table 5.1 summarize the
results from the discrete event simulation.
Table 5.1 Summary of simulation results for policy 1 and 2
Policy 1 Policy 2
Total overtime per day 10.6 8.4
Mean number of OTs with overtime per day 3.6 3.8
Mean emergency patient’s waiting time 74 (±4.4) 8 (±0.5)
OT utilization 74% 77%
From Table 5.1 it is clear that policy 2 outperforms policy 1 on all given out-
comes.
Under Policy 1, all emergency patients were operated on within 7 hours
with a mean waiting time of 74 (±4.4) min. Under Policy 2, all emergency patients
were operated upon within 80 min with a mean waiting time of 8 (±0.5) min. OT
utilization for policy 1 was 74% and 77% for policy 2. Policy 1 resulted in 10.6
hours of overtime on average per day and policy 2 resulted in 8.4. Policy 2, with
emergency capacity allocated to all elective OTs, thus substantially outperforms
policy 1, on all outcomes measured.
13
Fig. 5.5 Cumulative percentage of emergency patients in policy 1 and 2 (simulation results)
Table 5.2 summarizes the results of additional simulation experiments in which
we vary the number of emergency OTs (0, 1, 2, or 3) as well as the number of
elective OTs used for emergency surgeries (0, 5, 10, or 15). We use the case mix
of the previous experiment, but resize the problem to 15 elective OTs (instead of
12). Furthermore, approximately 10% of the surgeries are emergency surgeries.
The results show that policy 2 (dealing with emergencies in (some) elective OTs)
results in improved emergency waiting performances, at the expense of increased
waiting time of the elective surgeries. A mixed policy combines the advantages of
both policies – the table can be used as a guideline to make a trade-off.
Table 5.2 Simulation results: (1) average and (2) maximum emergency surgery waiting
time (minutes), (3) percentage of emergency surgeries that has to wait, (4) average elective
surgery waiting time (minutes)
Number of emergency OTs
Elective OTs
used for
emergency
0 1 2 3
0 -
21.9 2.4 0.5
3026 949 292
4.4% 2.4% 1.1%
12.6 12.6 12.6
5
1.3 0.6 0.3 0.1
204.9 152.7 113.1 83.3
4.5% 2.9% 1.5% 0.7%
32.3 21.2 14.2 11.4
10 0.5 0.3 0.1 0.1
0%
20%
40%
60%
80%
100%
10
40
70
100
130
160
190
220
250
280
310
340
370
400
Cumulative % of urgent surgery
Waiting time (minutes)
Policy 1
Policy 2
14
94.2 76.3 63.8 50.2
4.2% 2.6% 1.3% 0.6%
22.2 16.0 12.1 10.3
15
0.3 0.2 0.1 0.0
60.3 52.3 43.3 36.0
4.0% 2.5% 1.2% 0.5%
18.6 14.9 11.6 10.0
5.6 Tactical: designing a master surgical schedule to level ward
usage
Managers are inclined to solve problems at the moment they occur (i.e., on the op-
erational level). In Hans et al. (2011) we refer to this phenomenon as the “real-
time hype” of managers. For healthcare managers, while inundated with opera-
tional problems, the universal panacea for all productivity related problems is
“more capacity”. It is thereby often overlooked to tactically allocate and reorgan-
ize the available resources, which may turn out to be even more effective, and will
certainly be cheaper. However, due to its longer (intermediate) planning horizon,
tactical planning is less tangible and inherently more abstract than operational
planning. In the majority of our healthcare process optimization research projects
we find that the tactical planning level is typically not formalized and overlooked.
Tactical planning decisions are rather a result of historical development (“This
year’s tactical plan is last year’s tactical plan”), than a result of periodic planning.
We also find they have often evolved to hard constraints for operational planning
(“We don’t do orthopedic patients on Wednesday afternoons. Why? Well, we just
don’t!”).
This is also typical for the tactical planning of OTs, the block planning or
Master Surgical Scheduling (MSS) problem, which concerns the weekly allocation
of OT-days to surgical specialties (or surgeons). To a surgeon: “operating theatre
6 on Monday is her/his OT”. Re-allocating OT-days may however lead to a more
stable workload in subsequent departments (wards, ICU), and even reduce the re-
quired capacity of these departments. In this section we present a model to analyze
and improve the impact of the MSS on the resource usage in subsequent depart-
ments.
5.6.1 General problem description
Tactical OT planning, typically involves the assignment of OT capacity to aggre-
gate patient groups (i.e. patient cohorts) for a fixed planning horizon. This as-
signment should reflect the strategic goals of the hospital. For example, consider a
p
lanni
n
1000
o
special
t
geries
p
ning h
o
which
p
lishe
d
days d
u
to plan
ule als
o
how
m
when
c
ning h
o
availa
b
case
m
tients)
t
Q). Fo
r
assign
e
descri
b
block
MSS (
i
in Fig
u
that m
u
The M
S
LOS o
f
how p
a
n
g horizon o
f
o
rthopedic su
r
t
y should be
p
er month.
The tactica
l
o
rizon such
t
operational l
e
d
with a MSS.
u
ring the pla
n
their other f
u
o
allows the
O
m
any pieces o
f
c
an OT maint
e
When an
M
o
rizon are kn
o
b
le, etc. Wha
t
m
ix is expecte
d
t
o OT time is
A MSS rep
r
r
each day
e
d to one of
t
b
ed by the as
 where ∈
i
.e. before sp
e
u
re 5.6 where
u
ltiple blocks
Fig.
5
S
S is defined
f
any patien
t
.
a
tients overla
p
f
one month
a
r
ge
r
ies over t
h
assigned eno
u
l
plan is used
t
hat long te
r
m
e
vel planning
The MSS de
f
n
ning horizon
u
nctions, suc
h
O
T departme
n
f
equipment a
r
e
nance happe
n
M
SS is being
d
o
wn, i.e. whi
c
is known is
d
. Hence the
a
the primary
fa
r
esents a rep
e
1,2,,
t
he available
s
signment of
1,2,, a
n
e
cialties have
b
each cell rep
r
are assigned
t
5
.6 Example e
m
for period Q
a
Figure 5.7 d
i
p
.
a
nd a hospita
l
h
e next six
m
u
gh OT capa
c
to organize
c
m
goals are
m
can be base
d
f
ines which s
u
. Such a sch
e
h
as outpatien
t
n
t to make th
e
r
e needed ea
c
n
, etc.
developed, n
o
ch patients
w
that a certai
n
a
ssignment o
f
f
actor conside
r
e
titive pattern
in the MSS e
a
surgical spec
i
a surgical sp
n
d ∈1,2,
b
een assigne
d
r
esents an op
e
t
o a single sp
e
m
pty Master Su
r
and executed
i
splays how t
h
l
with a strat
e
m
onths, then t
h
c
ity to compl
e
c
apacity over
m
et and to c
r
d
. In the OT
t
u
rgical speci
a
e
dule allows t
h
t
clinics, edu
c
e
ir own plan
n
c
h day, what
a
o
t all of the
d
w
ill show up,
w
n
volume of
p
f
patient coho
r
ed when des
i
over a certai
n
a
ch of the I a
v
i
alties. More
ecialty j to
e
,. Using t
h
d
operating th
e
e
rating theatr
e
e
cialty on the
r
gery Schedule
repeatedly.
L
h
e multiple
M
e
gic goal to
c
h
e orthopedic
e
te 1000/6
an intermedi
a
r
eate a struct
u
t
his is usuall
y
l
ties operate
o
h
e surgical s
p
c
ation, etc. T
h
ing decisions
,
a
re the staffi
n
d
etails about
t
w
hich doctor
s
p
atients with
a
r
ts (not indiv
i
i
gning a MSS
.
n
number of
d
v
ailable OTs
h
p
recisely, th
e
e
ach operatin
g
h
is notation,
a
e
atre blocks)
i
e
block. It is
c
s
ame day.
(MSS)
et M
b
e the
m
M
SS cycles re
15
c
omplete
surgical
167 sur-
a
te plan-
u
re from
y
accom-
o
n which
p
ecialties
h
e sched-
, such as
n
g levels,
t
he plan-
s
will be
a certain
i
dual pa-
.
d
ays (say
h
as to be
e
MSS is
g
theatre
a
n empty
i
s shown
common
m
aximum
e
pea
t
and
16
In this
s
the mo
d
termedi
a
forman
c
MSS, c
o
inpatie
n
p
ropos
a
tic to fi
n
sion.
of the
M
distribu
t
termine
recover
i
ing inp
a
has the
p
ute th
e
for eac
h
dischar
g
experi
m
b
er of
e
distribu
t
Thus o
n
charges
tomorr
o
b
er of
a
each da
y
tion for
for the
d
Finally,
The for
m
the M
S
1,2,
,
gical s
p
crete di
Fig. 5.7 I
l
s
ection we d
e
d
el is to make
a
te term plan
n
c
e goals are
a
o
mpute a nu
m
n
ts. This ap
p
r
o
a
ls. Adopting
n
d the best
M
The aim is t
M
SS. We do t
h
t
ions. We th
e
the number
o
i
ng, we predi
c
a
tient care, di
s
Consider a
s
option of sta
y
e
probability
o
h
day of the
p
g
e probabiliti
e
m
ents have tw
o
e
xperiments
r
t
ion (assumi
n
n
any day an
d
and consequ
o
w. For a giv
e
a
dmission to
y
) for each p
a
the admissio
n
d
ischarge rat
e
using discre
t
m
al model de
Assume tha
t
S
S be define
d
,
indexes t
h
p
ecialty j
b
e c
h
stribution for
l
lustration of t
h
e
scribe a mo
d
a cyclical as
s
n
ing horizon,
a
chieved. W
e
m
ber of work
l
o
ach does not
t
his approac
h
M
SS proposal
o determine t
h
is by model
i
e
n add these
d
o
f patients re
c
c
t a number o
s
charges and
s
s
ingle patient
y
ing or bein
g
o
f being disc
h
p
atient’s LOS.
e
s. From pro
b
o
(and only t
w
esulting in e
a
n
g experimen
t
d
for each pat
i
e
ntly we kno
w
e
n MSS and
u
the ward (i.
e
a
tient cohort.
T
n
rate and usi
n
e
and we can
e
t
e convolutio
n
s
cription foll
o
t
each surgica
l
d
such that
h
e OTs and
h
aracterized
b
the number
o
h
e overlap bet
w
d
el by Vanbe
r
s
ignment of
O
such that str
a
e
allow for a
l
oad metrics
a
find the opti
m
h
to be used i
n
is of course
a
t
he number o
f
i
ng the recov
e
d
iscrete distri
b
c
overing. On
c
o
f workload
m
s
pecialized in
p
t
who is reco
v
g
discharged.
h
arged and c
o
.
Now consid
e
b
ability theo
r
w
o) outcome
s
a
ch outcome
t
s are indepe
n
i
ent cohort,
w
w
the numbe
r
u
sing historic
e
. the numbe
r
Thus for eac
h
n
g a binomial
e
asily compu
t
n
s we can co
m
o
ws.
l
specialty re
p
 is an o
p
1,2,,
b
y two param
o
f surgeries c
a
w
een multiple
M
r
kel et al. (20
O
T time to pat
i
a
tegic “produ
c
stochastic
L
a
ssociated wi
t
m
al MSS but
r
n
conjunction
a
natural and
f
patients in r
e
e
ring patient
c
b
utions (with
c
e we know t
h
m
etrics includi
n
p
atient care.
v
ering from
s
From histori
c
nversely the
p
e
r a cohort o
f
r
y it is know
n
s
, then the pr
o
can be comp
u
n
dent and id
e
w
e can comp
u
r
of patients
w
r
ecords, we c
r
of complet
e
h
patient coho
r
distribution
w
t
e the ward o
c
m
pute the ov
e
p
resents a sin
g
p
erating thea
t
the days in
a
eters and
a
rried out in
o
M
SS cycles
1
1). The obj
e
i
ent cohorts f
o
c
tion levels”
a
OS and, for
h
recovering
r
ather evalua
t
with a searc
h
very plausibl
e
e
covery as a
f
c
ohorts with
b
convolution
s
h
e number of
n
g admission
s
s
urgery and e
a
c
al data we c
a
p
robability o
f
f
patients wit
h
n
that when
m
o
bability for t
h
u
ted with a
b
e
ntically dist
r
t
e the numbe
r
w
ho will rem
a
a
n compu
t
e t
h
e
d inpatient s
u
r
t we have a
d
w
e have a dist
r
c
cupancy dist
r
e
rall ward occ
g
le patient co
h
t
re block w
h
a
cycle. Let e
a
, where
i
o
ne OT bloc
k
e
ctive of
o
r an in-
a
nd per-
a given
surgical
t
es MSS
h
heuris-
e exten-
f
unction
b
inomial
s
) to de-
patients
s, ongo-
ach day
a
n com-
f
staying
h
similar
m
ultiple
h
e num-
b
inomial
r
ibuted).
r
of dis-
a
in until
h
e num-
urgeries
d
istribu-
t
ribution
r
ibution.
c
upancy.
h
ort. Let
h
ere ∈
ach sur-
is a dis-
k
and
17
the probability that a patient, who is still in the ward on day n, is to be discharged
that day (0,1,,, where denotes the maximum LOS for specialty j).
Using and
as model inputs, for a given MSS the probability distri-
bution for the number of recovering patients on each day q can be computed. The
required number of beds is computed with the following three steps. Step 1 com-
putes the distribution of recovering patients from a single OT block of a specialty
j; i.e. we essentially pre-calculate the distribution of recovering patients expected
from an OT block of a specialty. In Step 2, we consider a given MSS and use the
result from Step 1 to compute the distribution of recovering patients given a single
cycle of the MSS. Finally in Step 3 we incorporate recurring MSSs and compute
the probability distribution of recovering patients on each day q.
Step 1: For each specialty j we use the binomial distribution to compute the num-
ber of beds required from the day of surgery 1 until . Since we know
the probability distribution for the number of patients having surgery (), which
equates to the number of beds needed on day 0, we can use the binomial dis-
tribution to iteratively compute the probability of needing beds on all days 0.
Formally, the distribution for the number of recovering patients on day n is recur-
sively computed by:
0



 1

.
Step 2: We calculate for each OT block  the impact this OT block has on the
number of recovering patients in the hospital on days q, q+1, …. If j denotes the
specialty assigned to OT block , then let
, be the distribution for the number
of recovering patients of OT block  on day 1,2,,,1,. It follows
that:
,


where 0 means
,01. Let be a discrete distribution for the total number
of recovering patients on day m resulting from a single MSS cycle. Since recover-
ing patients do not interfere with each other we can simply iteratively add the dis-
tributions of all the OT blocks corresponding to the day m to get . Adding two
independent discrete distributions is done by discrete convolutions which we indi-
cated by “”. For example, let A and B be two independent discrete distributions.
Then , which is computed by:


18
where τ is equal to the largest x value with a positive probability that can result
from ∗ (e.g. if the maximum value of A is 3 and the maximum value of B is 4,
then when convoluted the maximum value of the resulting distribution is 7, there-
fore in this example 7). Using this notation, is computed by:

,∗
,∗…∗
,∗
,∗…∗
,
Step 3: We now consider a series of MSSs to compute the steady-state probability
distribution of recovering patients. The cyclic structure of the MSS implies that
patients receiving surgery during one cycle may overlap with patients from the
next cycle. In the case of a small Q patients from many different cycles may over-
lap.
In Step 2 we have computed for a single MSS in isolation. Let M be the last
day where there is still a positive probability that a recovering patient is present in
. To calculate the overall distribution of recovering patients when the MSS is
repeatedly executed we must take into account / consecutive MSSs. Let 
denote the probability distribution of recovering patients on day q of the MSS cy-
cle, resulting from the consecutive MSSs. Since the MSS does not change from
cycle to cycle,  is the same for all MSS cycles. Such a result, where the proba-
bilities of various states remain constant over time, is referred to as a steady-state
result. Using discrete convolutions,  is computed by:
∗∗∗…∗/
From this result a number of workload metrics can be derived. To determine the
demand for ward beds from the variable  consider the following example. Let
the staffing policy of the hospital be such that they staff for the 90th percentile of
demand and let denote the 90th percentile of demand on day q. It follows that
is also the number of staffed beds needed on day q, and is computed from 
as follows: |0.9
In practice, patients tend to be segregated into different wards depending
on the type of surgery they received. To incorporate this segregation into the mod-
el and to consequently have recovering patient distributions for each ward, a mi-
nor modification needs to be made to the model. Let be the set of specialties j
whose patients are admitted to ward k. Then in Step 2 we only have to consider
those OT blocks assigned to a specialty in and continue with the calculations.
Ward occupancy alone does not fully account for the workload associated
with caring for recovering patients. During patient admissions and discharges the
nursing workload can increase. From the model the probability distribution for
daily admissions and discharges can be computed. To compute the admission rate,
19
set 1 for all j and repeat the steps above. The resulting  will denote the
admissions on day q.
The discharge rate is the rate at which patients leave the ward and can be
computed by adding an additional calculation in Step 1. Let
be a discrete dis-
tribution for the number of discharges from specialty j on day n which is comput-
ed as follows:
ℙ

 
1
ℙ

Finally, after computing
, one can set

and continue with Step 2. By
doing so, the resulting  will denote the distribution for daily discharges for
each day q of the MSS.
The inherent assumption of the described method is that all patients with
a patient cohort have equal probability of being discharged and that it is independ-
ent of other patients, i.e. it is assumed that patients are identically distributed and
independent. The independence assumption implies that the amount of time one
patient is in the hospital does not influence the amount of time another patient is in
the hospital. This seems like a natural assumption in most cases and appropriate so
long as surgeries are rarely cancelled due to a bed shortage (cancellations due to
bed shortages creates a dependency). The identically distributed requirement
means that we must compute the number of beds needed tomorrow (and the num-
ber of case completed in one OT block), for all identically distributed cohorts of
patients separately. In other words, the parameters of the binomial distribution
must reflect all of the patients in a given cohort.
5.6.2 General results
The model was applied at the Netherlands Cancer Institute - Antoni van Leeuwen-
hoek Hospital (NKI-AVL) to support the design of a new MSS. Selected results
from Vanberkel et al. (2011-2) are summarized in this subsection.
Management at NKI-AVL strive to staff enough beds such that for 90%
of the week days there is sufficient coverage. This implies that on 10% of the days
they will be required to call in additional staff. Using the model a number of MSS
proposals were evaluated and eventually staff choose an MSS that the model pre-
dicted would lead to a balanced ward occupancy.
An unbalanced ward occupancy makes staff scheduling, and ward opera-
tions, difficult. Early in the week, beds would be underutilized whereas later in the
week, beds would become highly utilized and the risk of a shortage would in-
crease. Such peaks and valleys represent variation in the system which possibly
could be eliminated with a different MSS. This variation leads to significant prob-
lems, particularly as the wards approach peak capacity. For example, when inpa-
20
tient w
a
ten scr
a
ditional
the elec
ed time
b
le, it
o
ward c
a
Either
w
care. A
l
out an
e
minimi
z
over a
3
used fo
r
fit tests
model.
cle) we
r
the sev
e
ward o
c
Fig. 5
.
5.6.3
D
The m
a
ward st
a
a soluti
o
several
ly predi
strictio
n
a
rds reach th
e
a
mble to try a
n
staff are call
e
tive surgery i
for surgeons
o
ften means a
a
pable of cari
n
w
ay, extra wo
r
l
though com
p
e
xorbitant am
o
z
e occurrence
After imple
m
3
3 week peri
o
r
each day of
, these obse
r
Six of the se
v
r
e found to b
e
e
nth day, this
c
cupancy wit
h
.
8 Comparison
D
iscussion
a
in benefit of
a
ff, thereby p
r
o
n. Staff was
modification
s
ct model out
p
In this proj
n
s as unchang
e
ir peak capac
n
d make a be
d
e
d in (or in r
a
s cancelled),
w
and extra an
x
patient was t
r
n
g for the pa
t
r
k is required
p
letely elimin
a
o
unt of resou
r
s
.
m
enting the
n
o
d. From thes
e
the MSS cyc
l
v
ed distribut
i
v
en projected
e
a good fit fo
was true at
a
h
the observe
d
of the projecte
d
using the m
o
r
oviding a pl
a
quick to em
b
s
to the origin
a
p
ut for a give
n
ect we treat
e
e
able. It is po
c
ity and a pat
i
d
available. I
f
a
re cases whe
n
w
hich causes
x
iety for pati
e
t
ransferred fr
o
t
ient but not
t
by ward staf
f
a
ting such pr
o
r
ces, sound p
n
ew MSS, t
h
e observatio
n
l
e were deriv
e
i
ons were co
m
distributions
o
r the observe
d
a
level α0.2
.
d
ward occup
a
d and observe
d
o
del was to b
a
tform from
w
b
race the mo
d
al MSS, at w
h
n
modificatio
n
e
d the equip
m
ssible that fu
r
i
ent admissio
n
f
one cannot b
n
additional
s
extra work f
o
e
nts. When a
b
o
m one ward
t
t
he designate
d
f
and there is
a
o
blems is lik
e
l
anning ahea
d
h
e ward occ
u
n
s, probability
e
d. Using Ch
i
m
pared to th
(one for eac
h
d
data at a le
v
.
Figure 5.8 c
a
ncy during th
d
(90th percent
i
e
able to qua
w
hich they co
u
d
el output, pa
r
h
ich point the
y
n
.
m
ent and p
h
r
ther improve
m
n
is pending,
s
e made avail
a
s
taff cannot b
e
o
r OR planne
r
b
ed was mad
e
t
o another (o
f
d
one) or dis
c
a
disruption i
n
e
ly not possib
d
of time ma
y
u
pancy was
o
distributions
i
-square good
n
o
se projecte
d
h
day of the
M
v
el α0.05,
w
o
mpare the p
r
e
33 week pe
r
le) ward occup
a
n
tify the con
c
u
ld begin to n
r
ticularly afte
r
y
were able t
o
h
ysician sche
d
m
ents in the
w
staff of-
a
ble, ad-
e
found,
r
s, wast-
e
availa-
f
ten to a
c
harged.
n
patient
b
le wi
t
h-
y
help to
o
bserved
of beds
d
ness-of-
d
by the
M
SS cy-
w
hile for
rojected
r
iod.
ancies
c
erns of
n
egotiate
r seeing
o
rough-
d
ule
r
e-
w
ard oc-
21
cupancy could have been achieved if these restrictions were relaxed. In this way
the model can be used to illustrate the benefits of buying an extra piece of equip-
ment or of changing physicians’ schedules. An additional restriction, which if re-
laxed may have allowed further improvements, is the assignment of wards to sur-
gical specialties. In other words, in addition to changing when a specialty
operates, it may prove advantageous to change which ward the patients are admit-
ted to after surgery. Finally, we chose the best MSS from those created through
swapping OR block and surgical specialty assignments. It is possible that a search
heuristic may have found a better MSS, although it would have required the many
surgical department restrictions to be modeled and the more complex model may
not have garnered the same level of staff understanding and support.
Oostrum et al. (2008-2) propose another approach, where the MSS is
planned in more detail: here it comprises of a cyclical schedule of frequently oc-
curring elective surgery types. The resulting combinatorial optimization problem
is to determine a MSS that balances OT utilization and ward occupancy. By
scheduling surgery types, the surgeon/surgical specialty can assign a patient’s
name at a later time, without affecting the performance of the MSS. The model
considers stochastic OT capacity constraints and empirical LOS distributions. As
the resulting problem is NP-hard, heuristics are provided. For a review on the suit-
ability and managerial implication of this particular MSS approach see Oostrum et
al. (2010).
5.7 Operational: elective surgery scheduling and sequencing
Operational planning and scheduling of operating theatres is arguably one of the
most popular topics in the healthcare operations research literature. The literature
reviews of Cardoen et al. (2010) and Guerriero and Guido (2011) outline many
contributions regarding the elective surgery scheduling and sequencing literature.
Cardoen et al. (2010-2) also propose a classification scheme for operating theatre
planning and scheduling problems, which contains four descriptive fields
|||. Here, holds the patient characteristics, the delineation of the deci-
sion, the extent to which uncertainty is incorporated, and the performance
measures.
In previous work (Hans et al. 2008) we demonstrated that by combining
advanced optimization techniques with extensive historical statistical records on
surgery durations, the OT department utilization can be improved significantly.
We demonstrated that, if slack time is reserved in OTs according to the method
described in Section 5.3.1 (particularly equation 5.6), the portfolio effect can be
exploited in a local search meta-heuristic as follows. By swapping surgeries be-
tween OTs (1-swap or 2-swap), the total slack time of both involved OTs is af-
fected. By clustering surgeries with similar duration variability characteristics, the
total slack time is reduced due to the portfolio effect. This principle can be used in
a local search heuristic to minimize the total slack time, and thus free OT time. A
22
result of the portfolio optimization is that the fragmentation of the free OT time is
minimized. In fact, OTs in resulting solutions are either filled to a great extent
with surgery and slack time, or are empty. As a result, OTs can be closed, or time
is freed to perform more surgeries.
In this section we discuss the optimization of the elective surgery sched-
ule, in order to minimize emergency surgery waiting time. This problem follows
from policy 2 outlined in Section 5.5 (i.e., emergency surgeries are dealt with in
elective OTs).
5.7.1 General problem description
Emergency surgery waiting time increases a patient’s risk of postopera-
tive complications and morbidity. When dealing with emergency patients accord-
ing to policy 2 (Section 5.5), waiting time will occur when all elective OTs are
busy. Typically, at the beginning of the regular working day all OTs will be busy
with long procedures, as surgeries are often scheduled according to the LPT rule.
As a result, emergency surgeries that arrive just after the start of the elective pro-
gram may have to wait a long time, as surgeries cannot be preempted. This pleads
for scheduling a short surgery at the beginning of the day, to obtain a so-called
“Break-in-Moment” (BIM), at an early time for emergency surgeries. Extending
on this idea, we may sequence the elective surgeries within their assigned OTs in
such a way, that their expected completion times, which are BIMs for emergency
surgery, are spread as evenly as possible. We do not re-assign surgeries to other
OTs, but instead only re-sequence elective surgeries within their assigned OT.
This is illustrated in Figure 5.9: the BIMs are clearly spread more evenly after re-
sequencing the surgeries.
OT1 OT2 OT3 OT1 OT2 OT3
Initialsolution AfterBIMoptimization
BIMs
Fig. 5.9 Example BIM optimization
23
The problem of sequencing elective surgeries in such a way that the BIMs are
spread as evenly as possible (or, alternatively, the break-in-intervals/BIIs are min-
imized) is in fact a new type of scheduling problem. This innovative idea was a re-
sult of a MSc thesis project at Erasmus MC (Lans et al. 2006), where it was prov-
en that the problem is NP-hard by reduction from 3-partition.
We assume, as illustrated in Figure 5.9, that surgeries are executed direct-
ly after another, i.e., there is no planned slack between surgeries. The planning
horizon is within a day, and starts on the first moment when all OTs are scheduled
to have elective surgeries. If all OTs start at the same time, then this time marks
the start of the planning horizon. It ends on the first moment when there is an OT
without a scheduled surgery, since after this moment there are infinitely many
BIMs. The objective is to lexicographically minimize the largest break-in-intervals
(BIIs). In other words, we minimize the largest BII, then the second largest (with-
out affecting the largest), etc. The reason that we do not only minimize the largest
BII is that the expected duration of the shortest surgery is a lower bound to the
longest BII. This can be seen as follows: assuming all OTs start at the same time,
placing the shortest surgery at the beginning of its OT gives a BII that cannot be
shortened.
In forthcoming work we will propose various exact and heuristic ap-
proaches for the BIM/BII optimization problem. Here we give the results of a
Simulated Annealing (SA) local search heuristic, which iteratively swaps surgeries
within their sequence. The SA method uses the following parameters: start tem-
perature 0.2, final temperature 0.0001, Markov chain length 150, decrease factor
0.8. We fix the shortest surgery at the beginning of its OT.
5.7.2 General results
We generate instances with the case mix of Section 5.5 (academic hospital Eras-
mus MC), scaled to fill 4, 8 or 12 OTs. Surgeries are scheduled “First Fit” (Hans
et al. 2008). First Fit assigns surgeries from the top of the list to the first available
OT plus an amount of slack (Section 5.3.1, equation 5.6) to achieve a 30% proba-
bility of overtime caused by surgery duration variability, until no surgery can be
found anymore that fits in the remaining OT capacity. Each instance has two vari-
ants: with full flexibility (all surgery sequences are allowed), and with reduced
flexibility (randomly,
 of the first surgeries are fixed on this position in their
OT, and
 of the last surgeries are fixed on this position in their OT). For exam-
ple, surgeries on children are typically done first, and surgeries after which exten-
sive OT cleaning is required are typically done last.
Table 5.3 presents the results for the SA algorithm. It compares the solu-
tions found by SA to the initial First Fit solution (which does not aim to optimize
BIM/BIIs). Particularly, it shows the frequency of the BIIs of size >15, >30, …,
24
>90 minutes. Each number is an average over 260 instances (52 weeks of 5 work-
ing days). SA solves each instance in less than 2 seconds. Clearly, the large inter-
vals are eliminated to a great extent, the more so when there are more OTs (and
thus more BIMs).
Table 5.3 Avg. frequency of break-in-interval size (initial solution SA solu-
tion; 260 instances per parameter setting)
#
OTs Reduced
flexibility >90 min. >75 min. >60 min. >45 min. >30 min. >15 min.
4 No
1.010.29 1.510.67 2.011.50 2.722.84 4.095.52 5.517.11
-71.3% -55.6% -25.4% 4.4% 35% 29%
4 Yes
1.010.30 1.510.74 2.001.59 2.722.85 4.105.47 5.557.01
-70.3% -51% -20.5% 4.8% 33.4% 26.3%
8 No
0.480.00 0.820.01 1.210.09 1.970.46 3.843.75 6.8910.22
-100% -98.8% -92.6% -76.6% -2.3% 48.3%
8 Yes
0.470.00 0.820.02 1.230.11 1.940.56 3.823.91 6.8810.02
-100% -97.6% -91.1% -71.1% 2.4% 45.6%
12 No
0.330.00 0.690.00 0.950.02 1.470.11 3.141.35 6.9710.49
-100% -100% -97.9% -92.5% -57% 50.5%
12 Yes
0.360.00 0.700.00 0.950.02 1.460.13 3.151.58 6.9210.26
-100% -100% -97.9% -91.1% -49.8% 48.3%
The question now is what impact these optimized BIMs/BIIs have on emergency
surgery waiting time, particularly given the fact that elective surgery durations are
stochastic, and the BIMs are expected surgery completion times. Table 5.4 pre-
sents the results of a Monte Carlo simulation of 260 instances with 12 OTs and
reduced sequencing flexibility. The elective surgeries are assumed to have a
lognormal distribution. The emergency surgeries arrive according to a Poisson
process (on average 5.1 arrivals per day), and are served on a FCFS basis. Elective
surgeries are not preempted.
Table 5.4 Waiting time for the 1st, 2nd and 3rd arriving emergency patients (12
OTs, run length 780 days, max. relative error 10%, min. confidence level
90%)
1st emergency
surgery 2nd emergency
surgery 3rd emergency
surgery
Waiting
time
(minutes)
Initial
solution SA
solution Initial
solution SA
solution Initial
solution
SA
solu-
tion
< 10 28.8% 48.6% 34.9% 44.9% 40.4% 46.2%
< 20 53.0% 75.8% 56.9% 73.6% 63.0% 69.8%
< 30 70.5% 90.9% 71.8% 87.2% 76.3% 86.7%
25
We observe that the BIM/BII optimization by SA, despite the reduced flexibility,
has a significant impact on emergency surgery waiting. For example, the relative
number of first emergency patients that wait at most 10 minutes increases by 69%
from 28.8% to 48.6%. The improvement decreases with every next arriving emer-
gency patient of the day. This may be expected, as these emergency patients in-
creasingly distort the original schedule.
5.7.3 Discussion
BIM/BII optimization has a big impact on emergency surgery waiting. More re-
search is required into exact solution approaches, and perhaps applications of
BII/BIM optimization in other sectors. For healthcare, it is easy to implement: it
only requires re-sequencing of elective surgeries. As a first step, managers are ad-
vised to plan the shortest surgery at the beginning of the regular working day.
5.8 Future Directions
The operating theatre department offers challenging planning and control prob-
lems on all hierarchical levels of control. While operational planning and control
has received a lot of attention from the OR/OM in healthcare research community,
tactical planning is less exposed, and research has had less of an impact in practice
due to its inherent complexity. In our experience, decision support software tools
mostly focus on the operational planning level, whereas tools for the tactical plan-
ning level are scarce and are too simplified or limited in scope to deal with tactical
decision making. Future research therefore has to focus on the tactical level, to a
greater extent. This raises opportunities to expand the scope beyond the operating
theatre department. From our survey of healthcare models that encompass multi-
ple departments we concluded that researchers often model hospitals in a way that
reflects the limited/departmental view of healthcare managers (Vanberkel et al.
2010). The research scope should particularly include the polyclinics, where new
patients are taken in, and the wards, which are typically managed to follow the OT
department but whose workloads may be leveled significantly by tactically opti-
mizing the OT’s master surgery schedule. Ultimately, we should aim to optimize
the entire patient care pathway.
5.9 References
Anthony RN (1965) Planning and control systems: a framework for analysis. Harvard Business
School Division of Research, Boston
26
Bhattacharyya T, Vrahas MS, Morrison SM, Kim E, Wiklund RA, Smith RM, Rubash HE, The
Value of the Dedicated Orthopaedic Trauma Operating Room. J Trauma 60:1336–1341
Blake JT, Donald J (2002) Using Integer Programming to Allocate Operating Room Time at
Mount Sinai Hospital. Interfaces 32(2):63–73
Brailsford SC, Harper PR, Patel B, Pitt M (2009) An analysis of the academic literature on simu-
lation and modelling in health care. Journal of Simulation 3:130–140
Cardoen B, Demeulemeester E, Beliën J (2010) Operating room planning and scheduling: A lit-
erature review. European Journal of Operational Research 201:921–932
Cardoen B, Demeulemeester E, Beliën J (2010-2) Operating room planning and scheduling prob-
lems: a classification scheme. International Journal of Health Management and Information
1(1):71–83
Denton BT, Miller A, Balasubramanian H, Huschka T (2010) Optimal Allocation of Surgery
Blocks to Operating Rooms Under Uncertainty. Operations Research 58(4):802–816
Denton BT, Viapiano J, Vogl A (2007) Stochastic Optimization of Surgery Sequencing and Start
Time Scheduling Decisions, Health Care Management Science 10(1):13–24
Dexter F, Macario A, Traub RD (1999) Which algorithm for scheduling add–on elective cases
maximizes operating room utilization? Anesthesiology 91:1491–1500
Dexter F (2011) Website: http://www.franklindexter.com/bibliography_TOC.htm
Dexter F, O’Neill L (2001) Weekend operating room on–call staffing requirements. AORN
Journal 74:666–671
Dexter F, Epstein RH, Traub RD, Xiao Y (2004) Making management decisions on the day of
surgery based on operating room efficiency and patient waiting times. Anesthesiology
101:1444–1453
Ferrand Y, Magazine M, Rao U (2010) Comparing two operating–room–allocation policies for
elective and emergency surgeries. Proceedings of the 2010 Winter Simulation Conference.
Johansson B, Jain S, Montoya–Torres J, Hugan J, Yücesan E (eds)
Guerriero F, Guido R (2011) Operational research in the management of the operating theatre: a
survey. Health Care Management Science 14:89–114
HFMA (2005) Achieving operating room efficiency through process integration. Technical re-
port., Health Care Financial Management Association
Hall R (2006) Patient Flow: Reducing Delay in Healthcare Delivery, Springer
Hans EW, Wullink G, Houdenhoven M van, Kazemier G (2008) Robust Surgery Loading. Euro-
pean Journal of Operational Research 185:1038–1050
Hans EW, Houdenhoven M van, Hulshof PJH (2011) A framework for health care planning and
control. Memorandum 1938, Department of Applied Mathematics, University of Twente, En-
schede
Harper PR and Shahani AK (2002) Modelling for the Planning and Management of Bed Capaci-
ties in Hospitals. Journal of the Operational Research Society 53(1):11–18
Houdenhoven M van, Hans EW, Klein J, Wullink G, Kazemier G (2007) A norm utilisation for
scarce hospital resources: Evidence from operating rooms in a Dutch university hospital.
Journal of Medical Systems, 31(4):231–236
Lans M van der, Hans EW, Hurink JL, Wullink G (2006) Anticipating Urgent Surgery in Operat-
ing Room Departments. BETA working paper WP-158, ISSN: 1386–9213, University of
Twente, 2006
McIntosh C, Dexter F, Epstein RH (2006) The impact of service specific staffing, case schedul-
ing, turnovers, and first–case starts on anesthesia group and operating room productivity: a
tutorial using data from an Australian hospital. Anesthesia & Analgesia 103:1499–1516
Oostrum JM van, Houdenhoven M van, Vrielink MMJ, Klein J, Hans EW, Klimek M, Wullink
G, Steyerberg EW, Kazemier G (2008) A Simulation Model for Determining the Optimal
Size of Emergency Teams on Call in the Operating Room at Night. Anesthesia & Analgesia
107:1655–1662
Oostrum JM van, Bredenhoff E, Hans EW (2010) Suitability and managerial implications of a
Master Surgical Scheduling approach. Annals of Operations Research 178(1):91–104
27
Oostrum JM van, Houdenhoven M van, Hurink JL, Hans EW, Wullink G, Kazemier G (2008-2)
A master surgical scheduling approach for cyclic scheduling in operating room departments.
OR Spectrum, 30(2):355–374
Vanberkel PT, Boucherie RJ, Hans EW, Hurink JL, Lent WAM van, Harten WH van (2011) An
exact approach for relating recovering surgical patient workload to the master surgical sched-
ule. Journal of the Operational Research Society (forthcoming)
Vanberkel PT, Boucherie RJ, Hans EW, Hurink JL, Lent WAM van, Harten WH van (2011-2)
Accounting for Inpatient Wards when developing Master Surgical Schedules. Anesthesia &
Analgesia (forthcoming)
Vanberkel PT, Blake JT (2007) A Comprehensive Simulation for Wait Time Reduction and Ca-
pacity Planning Applied in General Surgery. Health Care Management Science, 10 (4) 373–
385, Special Issue on Simulation in Health Care. Anderson JG, Merode GG van (eds)
Vanberkel PT, Boucherie RJ, Hans EW, Hurink JL, Litvak N (2010) A Survey of Health Care
Models that Encompass Multiple Departments. International Journal of Health Management
and Information 1(1):37–69
Vissers JMH, Bertrand JWM, De Vries G (2001) A framework for production control in
healthcare organizations. Production Planning & Control 12:591–604
Wachtel RE, Dexter F (2008) Tactical increases in operating room block time for capacity plan-
ning should not be based on utilization. Anesthesia & Analgesia 106(1):215–22
Wixted JJ, Reed M, Eskander MS, Millar B, Anderson RC, Bagchi K, Kaur S, Franklin P, Le-
clair W (2008) The Effect of an Orthopedic Trauma Room on After–Hours Surgery at a Lev-
el One Trauma Center. J Orthop Trauma 22(4):234–236
Wullink G, Houdenhoven M van, Hans EW, Oostrum JM van, Lans M van der, Kazemier G
(2007) Closing emergency operating rooms improves efficiency. Journal of Medical Systems
31(6):543–546
Zijm WHM (2000) Towards intelligent manufacturing planning and control systems. OR Spec-
trum 22(3):313–345
... Tactical level planning is also termed master surgery scheduling, and it involves the division of allocated time for different surgical specialities and prepares a master surgery schedule (MSS). The operational level planning is also termed short-term planning, where a daily plan of the operating room is made, and patients are sequenced within the allocated operating rooms [8], [9]. Operational level planning consists of assigning an operating room to surgery patients, referred to as advanced scheduling and sequencing of patients in the assigned operating rooms, termed allocation scheduling. ...
... At the operational level planning daily plan of the operating room is made, and patients are allocated to operating rooms and sequenced within the allocated operating rooms [8], [9]. At an operational level, advanced scheduling includes assigning a date and an operating room, and allocation scheduling includes sequencing patients within the allocated operating rooms. ...
... Equation (8) states that each surgeon can be assigned to more than one operating room on a day d of planning horizon τ . Equation (9) ensures that the total number of scheduled patients is less than or equal to the demand of surgery patients in a planning horizon. Equations (10) and (11) put an upper bound on the no. of beds required by surgical patients in the ICU and ward, respectively. ...
Article
Full-text available
Operating room planning and scheduling significantly affect all hospital areas, including the intensive care unit and downstream wards. Planning and scheduling operating rooms integrated with intensive care units and downstream wards can lead to more stable plans and schedules that are less prone to cancellations. Thus, this study considers the operating room’s capacity and downstream units while making surgery-related decisions. A mixed integer linear programming model for integrated planning consisting of two stages is proposed. The first stage model maximizes the scheduled surgical time of all operating rooms. In contrast, the second stage model aims to minimize the makespan of patients in operating rooms by incorporating sequence-dependent setup time and the capacity constraints of all resources under consideration at both stages. A two-stage genetic artificial bee colony algorithm (TGABC) hybrid of genetic algorithm and artificial bee colony algorithm is proposed to solve the model. Taguchi design of experiments is employed to fine-tune the parameters of the proposed TGABC algorithm. Experiments are designed to evaluate the performance of the proposed TGABC algorithm with generated instances mimicking real data for different-sized problems, and results are presented. The proposed method is compared with the exact method and three standard metaheuristics. It provides near-optimal results in comparatively shorter CPU time. Moreover, it outperforms the genetic algorithm, artificial bee colony algorithm, and simulated annealing in terms of solution quality compared on the considered test instances.
... Over the last few decades, outpatient appointment scheduling has piqued the interest of various academics and clinicians, starting with the work of Welch and Bailey [13]. In recent decades, the literature has properly covered appointment scheduling, including outpatient scheduling [14], operation room scheduling [15][16][17][18][19][20][21][22][23][24], and medical examination scheduling [25][26][27][28][29]. Te primary issues with appointment scheduling are connected to optimizing healthcare resources through better usage of human resources and medical equipment, which results in shorter patient wait times. Numerous research studies have found that patient dissatisfaction with outpatient scheduling is typically due to long wait times and that reasonable wait times are necessarily based on clinical competency [30]. ...
... Te fndings indicate that there was reduced idle time between surgical patients as a result of the operation schedules as well as more operating room utilization with little overtime. Additionally, [20] created a unique two-stage stochastic mixed-integer-programming approach to handle surgical schedules across many operating rooms under uncertainty. Te major point of this study is that the availability of several operating rooms and assistance and support from other surgeons, particularly the chief staf surgeon, allows many procedures to be done concurrently. ...
... It regulates the amount of time on nonemergency patients. Restriction (20) guarantees that each patient who has surgery is admitted to the critical care unit. Te restriction above (21) guarantees that each patient receives a bed in an intensive care unit (ICU) for the number of days necessary for admission. ...
Article
Full-text available
Operating room scheduling is a prominent study topic due to its complexity and significance. The increasing number of technical operating room scheduling articles produced each year calls for another evaluation of the literature to enable academics to respond to new trends more quickly. The mathematical application of a model for the patient admission scheduling issue with stochastic arrivals and departures is the subject of this study. The approach for applying our model to real-world issues is discussed here. We present a solution technique for efficient computing, a numerical model analysis, and examples to demonstrate the methodology. This study looked at the challenge of assigning procedures to operate rooms in the face of ambiguity regarding surgery length and the arrival of emergency patients based on a flexible policy (capacity reservation). We demonstrate that the proposed methods derived from deterministic models are inadequate compared to the answers produced from our stochastic model using simple numerical examples. We also use heuristics to estimate the objective function to build more complicated numerical examples for large-scale issues, demonstrating that our methodology can be applied quickly to real-world situations that often include big information sets.
... We argue that the surgical process data from the OR benchmarking initiative of German hospitals especially has the 1 As of 2022. potential for detailed modeling approaches of the short-term ("operational") [37] OR planning, i.e., surgery scheduling, in particular. However, it can also be used for studies on OR process design. ...
... With detailed process data, overlapping processes can be modeled, which is, for example, directly being used by the research on OR process design [5,12,40]. A high level of process detail is generally not necessary for other typical OR planning problems on the strategic or tactical level, such as dimensioning and allocating OR resources [37]. Detailed process data could, however, be used for the strategic problem of layout planning [63], where the focus lies on the pathways of the different stakeholders in the OR. ...
Article
Full-text available
We present a freely available data set of surgical case mixes and surgery process duration distributions based on processed data from the German Operating Room Benchmarking initiative. This initiative collects surgical process data from over 320 German, Austrian, and Swiss hospitals. The data exhibits high levels of quantity, quality, standardization, and multi-dimensionality, making it especially valuable for operating room planning in Operations Research. We consider detailed steps of the perioperative process and group the data with respect to the hospital’s level of care, the surgery specialty, and the type of surgery patient. We compare case mixes for different subgroups and conclude that they differ significantly, demonstrating that it is necessary to test operating room planning methods in different settings, e.g., using data sets like ours. Further, we discuss limitations and future research directions. Finally, we encourage the extension and foundation of new operating room benchmarking initiatives and their usage for operating room planning.
... The demand quantities of medical devices can be derived from the surgery schedule as a result of operational surgery planning, which usually includes a planning horizon of one week, cf. [15]. The aim of operational surgery planning is to assign patients to a specific day of the week with a start time for the operation as well as an operating room and team. ...
... The space requirement vol k for reprocessing one unit of medical device k is uniformly distributed in the interval vol k ∈ [5,15]. The average space requirement vol κ is determined for each class κ. ...
Article
Full-text available
We present a new model formulation for a multiproduct dynamic order quantity problem with product returns and a reprocessing option. The optimization considers the limited shelf life of sterile medical devices as well as the capacity constraints of reprocessing and sterilization resources. The time-varying demand is known in advance and must be satisfied by purchasing new medical devices or by reprocessing used and expired devices. The objective is to determine a feasible procurement and reprocessing plan that minimizes the incurred costs. The problem is solved in a heuristic manner in two steps. First, we use a Dantzig-Wolfe reformulation of the underlying problem, and a column generation approach is applied to tighten the lower bound. In the next step, the obtained lower bound is transformed into a feasible solution using CPLEX. Our numerical results illustrate the high solution quality of this approach. The comparison with a simulation based on the first-come-first-served principle shows the advantage of integrated planning.
... Online operational decisions (1-day horizon) are real-time decisions made during the execution of the schedule. They deal with disruptions that deviate the performed schedule from the initial schedule (Hulshof et al., 2012;Hans and Vanberkel, 2012). Consequently, (Hulshof et al., 2012) defines online operational planning as "the control mechanisms that deal with monitoring the process and reacting to unplanned events". ...
... In other words, the OnOM is responsible for reducing the gaps between the initial schedule (IS) and the performed schedule (PS) by limiting or even preventing the impact of uncertainties on the ongoing surgical schedule (Hulshof et al., 2012;Hans and Vanberkel, 2012). As stated in Dexter et al. (2004), disruptions are inherent in the functioning of the surgical suite and its ORs. ...
... Some articles investigate decision-making regarding operating room planning and scheduling in hospitals at four hierarchical decision levels: strategic, tactical, offline operational, and online operational. The offline operational level deals with assigning patients to dates and sequencing the patients in the operating rooms and involves monitoring and controlling the schedule execution [3,4]. ...
Article
Full-text available
Planning and scheduling critical resources in hospitals is significant for better service and profit generation. The current research investigates an integrated planning and scheduling problem at different levels of operating rooms, intensive care units, and wards. The theory of constraints is applied to make plans and schedules for operating rooms based on the capacity constraints of the operating room itself and downstream wards. A mixed integer linear programming model is developed considering shifting bottleneck resources among the operating room, intensive care unit, and hospital wards to maximize the utilization of resources at all levels of planning. Different sizes of planning and scheduling problems of the hospital, including small, medium, and large sizes, are created with variable arrivals and surgery durations and solved using a CPLEX solver for validating the developed models. Later, the application of the proposed models in the real world to develop planning systems for hospitals is discussed, and future extensions are suggested.
... However, the presence of randomness and unpredictability inherent to surgical processes adds complexity and creates inefficiency to the OR system [23,32]. According to the time frame, OR planning and scheduling can be classified into four different hierarchical decision levels: strategic, tactical, offline operational, and online operational [17]. A large body of literature tries to deal with the uncertainty in the OR system by using various means at these decision levels. ...
Article
Full-text available
In hospitals, the efficient planning of the operating rooms (ORs) is difficult due to the uncertainty inherent to surgical services. This is especially true for the inpatient surgical department where complex and long surgeries are often performed along with surgeries on emergency patients. This paper aims to improve the scheduling of the inpatient department by partitioning the elective surgeries into the more predictable surgeries (MPS) group and the less predictable surgeries (LPS) group, based on surgery duration variability, and by scheduling each of the two surgery groups in different ORs. Through a simulation study that comprehensively investigates the impact of the partitioning on different performance measures under various environmental settings, we report important findings and insights. First, partitioning can effectively shorten the waiting times of elective patients for both MPS and LPS groups, but the option should be allowed to reassign patients from the MPS or LPS ORs to the other ORs when needed. Meanwhile, partitioning sometimes slightly increases the elective cancellation rate. Second, the ability to use the available capacity of the ORs as much as possible is key to reducing elective waiting times. Third, partitioning might slightly worsen the waiting times of emergency patients, while the slightly negative impact on emergency patients decreases when the number of ORs is higher. Fourth, the beneficial impact of partitioning on elective patients increases with an increased patient demand. Last, for the settings considered in this study there was no benefit in partitioning the elective patients into more than two groups.
Chapter
Operation theatre is one of the largest revenue centres of hospitals but also required high costs, therefore, the schedule of operation theatre should be effective and maximize the usage of operation theatre to improve their performance and meet patients’ satisfaction levels. This study is aimed to construct a one-month schedule of an operation theatre at the strategic level by using the Integer Linear Programming method, maximize the usage of operation theatres and compare the existing schedule with the constructed schedule to validate the effectiveness of the proposed model. The coding is programmed into AMPL software and the schedule is solved weekly since the demand of operating hours of each department is updated each week by adding the unfulfilled demand in the previous week to the original demand of the same department in the current week. The constructed schedule has allocated a total of 54 operation theatres in a month. It has also achieved 96.52% of the overall allocated operating hours compared to the existing schedule. The usage of operation theatre has been maximized and the effectiveness of the proposed model is validated. However, this study can be improved by using various optimization methods and taking more variables and parameters into account.KeywordsScheduling ProblemOperation theatreStrategic level
Article
Full-text available
This article describes a multi-dimensional approach to the classification of the research literature on simulation and modelling in health care. The aim of the study was to analyse the relative frequency of use of a range of operational research modelling approaches in health care, along with the specific domains of application and the level of implementation. Given the vast scale of the health care modelling literature, a novel review methodology was adopted, similar in concept to the approach of stratified sampling. The results provide new insights into the level of activity across many areas of application, highlighting important relationships and pointing to key areas of omission and neglect in the literature. In addition, the approach presented in this article provides a systematic and generic methodology that can be extended to other application domains as well as other types of information source in health-care modelling.
Article
Full-text available
Operating Room (OR) departments need to create robust surgical schedules that anticipate urgent surgery, while minimizing urgent surgery waiting time and overtime, and maximizing utilization. We consider two levels of planning and control to anticipate urgent surgery. At the tactical level, we study the allocation of slack for urgent surgery to one or more operating rooms, and at operational off-line level, we experiment with the sequencing of elective surgeries in the operating rooms to which slack for urgent surgery is allocated. We try to sequence the elective surgeries such that their completion times, which are break-in-moments (BIMs) for urgent surgery, are spread as equally as possible over the day. We refer to this problem as BIM optimization problem, which is NP-hard in the strong sense. In this paper, we develop and test various heuristics for this sequencing problem. By means of a simulation study, we compare five methods of anticipating urgent surgery: (1) concentrating slack for urgent surgery in a dedicated operating room, (2) allocating slack for urgent surgery to a subset of the operating rooms without BIM optimization and (3) with BIM optimization, and (4) allocating slack for urgent surgery to all operating rooms without BIM optimization, and (5) with BIM optimization. For the test instances, the computational experiments show that urgent surgery can be anticipated best by allocating slack for urgent surgery to all available operating rooms, and thus allowing urgent surgeries to interfere with the schedule of elective surgeries. Further savings in urgent surgery waiting time can be achieved by BIM optimization, especially for the first urgent surgical cases that arrive during a day.
Chapter
Full-text available
Rising expenditures spur healthcare organizations to organize their processes more efficiently and effectively. Unfortunately, healthcare planning and control lags behind manufacturing planning and control. We analyze existing planning and control concepts or frameworks for healthcare operations management and find that they do not address various important planning and control problems. We conclude that they only focus on hospitals and are too narrow, focusing on a single managerial area, such as resource capacity planning, or ignoring hierarchical levels. We propose a modern framework for healthcare planning and control that integrates all managerial areas in healthcare delivery operations and all hierarchical levels of control, to ensure completeness and coherence of responsibilities for every managerial area. The framework can be used to structure the various planning and control functions and their interaction. It is applicable to an individual department, an entire healthcare organization, and to a complete supply chain of cure and care providers. The framework can be used to identify and position various types of managerial problems, to demarcate the scope of organization interventions and to facilitate a dialogue between clinical staff and managers.
Article
Full-text available
The increasing interest in the domain of operating room planning and scheduling leads to a proliferation of problem types. The statement and the scope of the particular problems, however, are often unclear. In this paper, we report on a scheme to classify operating room planning and scheduling problems using multiple fields. Each field describes a specific set of characteristics of the particular problem by means of parameters, elements and optional further specifications. We also elaborate on the use of delimiters to separate the entries in the classification notation. Next to the formulation of the scheme, we examine its applicability on a range of problems that are encountered in recent literature. With the development of the classification scheme, we hope to structure and to clarify forthcoming research in this domain.
Article
Full-text available
The increasing interest in the domain of operating room planning and scheduling leads to a proliferation of problem types. The statement and the scope of the particular problems, however, are often unclear. In this paper, we report on a scheme to classify operating room planning and scheduling problems using multiple fields. Each field describes a specific set of characteristics of the particular problem by means of parameters, elements and optional further specifications. We also elaborate on the use of delimiters to separate the entries in the classification notation. Next to the formulation of the scheme, we examine its applicability on a range of problems that are encountered in recent literature. With the development of the classification scheme, we hope to structure and to clarify forthcoming research in this domain.
Article
The internal dynamics of a hospital represent a complex non-linear structure. Planning and management of bed capacities must be evaluated within an environment of uncertainty, variability and limited resources. A common approach is to plan and manage capacities based on simple deterministic spreadsheet calculations. This paper demonstrates that these calculations typically do not provide the appropriate information and result in underestimating true bed requirements. More sophisticated, flexible and necessarily detailed capacity models are needed. The development and use of such a simulation model is presented in this paper. The modelling work, in conjunction with a major UK NHS Trust, considers various types of patient flows, at the individual patient level, and resulting bed needs over time. The consequence of changes in capacity planning policies and management of existing capacities can be readily examined. The work has highlighted the need for evaluating hospital bed capacities in light of both bed occupancies and refused admission rates. The relationship between occupancy and refusals is complex and often overlooked by hospital managers.
Book
Patient Flow: Reducing Delay in Healthcare Delivery is dedicated to improving healthcare through reducing the delays experienced by patients. One aspect of this goal is to improve the flow of patients, so that they do not experience unnecessary waits as they flow through a healthcare system. Another aspect is ensuring that services are closely synchronized with patterns of patient demand. Still another aspect is ensuring that ancillary services, such as housekeeping and transportation, are fully coordinate with direct patient care. It is the first book treatment to have reduction in patient delay as its sole focus, and therefore, provides the foundation by which hospitals can implement change. Reflecting the highly interdisciplinary and practitioner nature of this book, the chapters have been written by doctors, nurses, industrial engineers, system engineers and geographers, and thus, these perspectives provide the comprehensive view needed to address the problem of patient delay.
Article
This article describes a multi-dimensional approach to the classification of the research literature on simulation and modelling in health care. The aim of the study was to analyse the relative frequency of use of a range of operational research modelling approaches in health care, along with the specific domains of application and the level of implementation. Given the vast scale of the health care modelling literature, a novel review methodology was adopted, similar in concept to the approach of stratified sampling. The results provide new insights into the level of activity across many areas of application, highlighting important relationships and pointing to key areas of omission and neglect in the literature. In addition, the approach presented in this article provides a systematic and generic methodology that can be extended to other application domains as well as other types of information source in health-care modelling.
Article
The paper presents a hierarchical framework for production control of hospitals which deals with the balance between service and efficiency, at all levels of planning and control. The framework is based on an analysis of the design requirements for hospital production control systems. These design requirements are translated into the control functions at different levels of planning required for hospital production control. The framework consists of five levels of planning and control: patient planning and control, patient group planning and control, resources planning and control, patient volumes planning and control and strategic planning, though this last level does not make part of production control as such. Each of the levels of the framework is further elaborated in terms of the decisions made regarding patient flows and resources, and the co-ordination of the different planning levels. Implications of the framework are discussed by describing some points where current practice deviates from assumptions made in our approach. Recommendations for future research and development of the planning framework are formulated.