You are on page 1of 14

1

A comparison of OTN and MPLS networks under


trafc uncertainty
Pietro Belotti, Kireeti Kompella, and Lloyd Noronha
AbstractWe study a special case of the problem of installing
capacity on backbone networks in the presence of an uncertain
trafc matrix. The uncertainty model is, in general, dened by
a set of linear constraints on the trafc demands. Owing to the
availability of statistics on the nominal (average) trafc matrix,
in our experiments each trafc demand is characterized by an
average and a peak volume trafc. To allow for trafc variability,
we assume that, throughout network operation, any subset of at
most K demands, whose elements can change over time, are
allowed to take on a peak value.
The problem consists in allocating sufcient capacity on the
network so as to route all trafc matrices included in the
uncertainty set. Network links can be equipped with different
technologies: optical transport network (OTN) switches and multi-
protocol label switching (MPLS) routers. These can be combined
in several ways to create: an OTN-only network, i.e., with OTN
switches only; an MPLS-only network; a combined network with
an OTN and an MPLS layer; and a bypass network, where trafc
enters and leaves the core of the network using MPLS routers, but
all the interim transit nodes are OTN switches. While the cost of
all these types of network is inuenced by uncertainty, a network
using MPLS routers can take advantage of statistical multiplexing
and reserve capacity by considering both the average and peak
values, thus limiting the total capacity to be installed.
When only one type of router is considered, the design problem
can be trivially solved without resorting to optimization methods.
A more general problem, where more than one technology can
be used to route the matrix, is modeled by Robust Optimization,
a paradigm yielding a solution that guarantees to work in
any scenario within the uncertainty set. Our computational
experiments prove that, for a realistic cost differential between
MPLS and OTN ports, the former yields the minimum-cost
network by taking advantage of statistical multiplexing while
supporting all possible demand combinations.
Index TermsStatistical multiplexing, robust optimization,
trafc uncertainty, MPLS, OTN.
I. MOTIVATION
Provisioning backbone telecommunication networks is a key
task in the design of the communication infrastructure. Net-
work operators have to cope with an ever-changing technology
market, which makes it difcult to choose the most appropriate
and cost-effective technology. One of the most important
factors is uncertainty in the trafc demand, which introduces
the tradeoff between an inexpensive circuit-switched network
(at the risk of having to drop demands due to insufcient
capacity) and a more conservative and robust network that
manages well any statistical variation, albeit more costly.
P. Belotti is with the Deptartment of Mathematical Sciences, Clemson
University, Clemson, SC 29630. Email: pbelott@clemson.edu
Kireeti Kompella and Lloyd Noronha are with Juniper Networks, Sunny-
vale, CA 94089. Email: {kireeti,noronha}@juniper.net
In this article, we compare two routing technologies, MPLS
and OTN, in terms of the total network cost, using a realistic
backbone network. Due to the high variability of the trafc
matrix (TM) or, equivalently, the fact that it is not known a
priori we target all of our experiments at a set of TMs.
The variability and unpredictability of TMs, which may
occur even on an hourly basis, is best tackled by routing
technologies that can take advantage of statistical multiplexing,
such as MPLS, which however have a higher per-port cost.
We show that several scenarios of trafc variability favor
MPLS-based networks, which are more efcient at accom-
modating volatile demands. Our claims are supported by an
optimization model that determines the minimum-cost network
as a combination of MPLS routers and OTN switches that
satises a set of TMs. We outline below the context of our
work and the main challenges we have addressed.
A. Optimization in network design
Network design problems can be broadly dened as follows:
the network is represented by a graph G = (V, E), where V is
the set of nodes and E is the set of links. A set Q of demands,
each represented by a triplet (s
q
, t
q
, v
q
), for q Q, is given.
Each triplet denes a source node s
q
, a destination node t
q
,
and a trafc volume v
q
, expressed in Mbps. The problem is to
determine the capacity to be installed on each link or node of
G so as to route all trafc demands while minimizing the total
installation cost, which is a function of the capacity installed
both on links and on nodes.
Various problems in the design of telecommunication net-
works can be formulated as optimization problems [1, 2, 3, 4].
Optimization models describe concisely and completely all
possible congurations of a network, and, when solved, pro-
vide the one with the best performanceintended here as a
generic term that can mean throughput, total cost, number of
demands routed, congestion, etc.
Optimization models consist of a set of variables, a set of
constraints, and an objective function. Variables can be contin-
uous when modeling ow or portions of demand, integer (e.g.
when modeling number of ports, bers, etc.), or binary, suited
for yes/no decisions. The value of these variables at the end
of the optimization process denes the optimal conguration
of the network, if any exists. The constraints identify a set of
congurations of the system, and have to be specied in such
a way that all and only possible congurations of the network
are considered. The objective function denes the criterion to
be used for choosing the best among the (often innitely many)
congurations. State-of-the-art solvers for several classes of
2
optimization problems are available; depending on the size
(number of variables/constraints) and type of problem (e.g.,
integer variables or nonlinear constraints), the solution times
can vary signicantly, hence special care is required to limit
the models complexity in the modeling phase.
In the remainder of the paper, we consider a network
optimization problem under the following assumptions:
1) All routing paths must be set up according to the OSPF
(Open Shortest Path First) protocol: given a function
w : E Z
+
dening the weight of each link, the
routing path of a demand q from source s
q
to destination
t
q
must be one of the shortest paths between s
q
and
t
q
according to the metric dened by the weights w
ij
,
which are known in advance. Further constraints on the
routing proportions, such as those required by Equal
Cost Multi-Path (ECMP), can also be considered.
2) The trafc matrix is unknown a priori and changes over
time, but varies within a bounded, non-empty set that is
known in advance.
3) The capacity to be installed corresponds to the number
of ports of different capacity (1.25, 2.5, and 10 Gbps,
for instance) at each node.
The rst assumption greatly simplies the network design
problem, and makes our models viable for relatively large
networks. In fact, if the weights w
ij
, for ij E, are given as
input, then the routing paths of all demands can be determined
by shortest path computations, which take negligible time even
in large networks. Thus, in most cases, the capacity allocation
can be done automatically and the set of feasible solutions can
be signicantly reduced. The last two assumptions, however,
require a more detailed description: assumption 2 is explained
in the remainder of this section, while the technologies we
have considered are detailed in Section II.
B. Network optimization under trafc uncertainty
Real-world network design problems are subject to uncer-
tainty in several contexts: cost, capacity, and reliability of link
and node equipment, and, most importantly, trafc matrix,
which is the focus of this paper. Although the set of source-
destination pairs may be known, the volume of trafc of each
demand is seldom known with accuracy. This happens for
several reasons: inaccurate measurements, dynamic behavior
of trafc, network failures inducing a shift in the demand,
to name only a few. The sources of trafc uncertainty are
multiple and need to be identied and accurately described:
for instance, it helps to know whether there are lower/upper
bounds on the volume of a demand from a source s to a
destination t of the network, bounds on the total trafc volume
or, again, statistical information on the TM.
From a modeling standpoint, the more accurate the approx-
imate trafc matrix, the better: a network that accommodates
a loose uncertainty set can be overly conservative, hence
very expensive. From a more computational point of view,
solving a network provisioning problem while taking into
account uncertainty in the problems parameters may render
the problem itself intractable, as the size of the associated
optimization problem (number of variables and constraints)
may increase substantially.
A common way to model this problem is Robust Optimiza-
tion [5]: the trafc demand is not known a priori, but a set
S of possible trafc matrices is given. The capacity installed
and the routing paths then must accommodate, at any time, all
trafc matrices in S. There are other approaches for solving
optimization problems under uncertainty, the most important
being Stochastic Programming [6], in which a probability dis-
tribution is associated with each element of the uncertainty set
[7, 8, 9, 10, 11]. Although Stochastic Programming can exploit
statistical information on uncertain parameters, it produces
even more difcult optimization problems. As a consequence,
depending on the size of the network, this modeling framework
can be impractical.
We use robust optimization for modeling our network design
problem. From now on, we consider the problem of installing
capacity on the network nodes in order to accommodate all
trafc matrices in a given set S, known a priori.
C. Previous work
The application of robust optimization to network design
under trafc uncertainty has been a hot topic in the last decade.
Dufeld et al. [12] introduce the hose model, an uncertainty set
where an upper bound for the incoming and outgoing demand
is dened for each node, i.e., the uncertainty set S is the
following set of vectors (v
q
)
qQ
:
S = {v R
|Q|
+
:

qQ:sq=i
v
q
b

i
i V,

qQ:dq=i
v
q
b

i
i V }.
Gupta et al. [13] study a particular case of this problem where
the capacity must be installed so that the edges with nonzero
capacity form a spanning tree of the original network topology.
They prove that this version of the problem can be solved
in polynomial time and propose an algorithm to solve the
problem (see also [14, 15]).
Ben Ameur and Kerivin [16] propose an uncertainty model
where the set S of trafc matrices is a non-empty, bounded
polyhedron in |Q| dimensions (|Q| is the number of trafc de-
mands) dened by a set of linear inequalities. This polyhedral
uncertainty model encloses the hose model as a special case.
This leads to a semi-innite linear programming (LP) problem,
an optimization problem with nitely many variables and
innitely many constraints: given that there is one constraint
for each trafc matrix, which in turn is a point of a non-empty,
bounded polyhedron, the number of constraints is innite.
Rather than solving the semi-innite LP explicitly, they
describe an iterative method that considers an initial set
of demands. At each step, after solving the current LP, a
separation problem is solved to generate a trafc demand that
cannot be routed. This demand is used to increase the total
capacity, and the algorithm continues until separation gives
no new violated demands, proving that all demands in the
polyhedron are satised.
Ben Tal and Nemirovski [17] introduce a class of robust
optimization problems where the uncertainty set is identied
3
by a vector of mean values and a covariance matrix, which
dene a condence ellipsoid that includes a safe portion
of all vectors of uncertain parameters. Optimizing over this
uncertain set amounts to solving a second order conic opti-
mization problem [18]. Although more difcult than Linear
Programming problems, second order conic problems are
convex optimization problems and therefore admit an efcient
solution method.
Bertsimas and Sim [19] introduce an uncertainty model
where all parameters are allowed to take on a lower or an upper
value, while imposing a maximum number K of demands that
are at the upper value simultaneously. This is a compact yet
meaningful way to express a large uncertainty set.
Belotti et al. [20] deal with a network design problem using
the aforementioned uncertainty set in the very special case
where K = 1. At any point in time, all demands are at their
lower value except for one, which is at its upper value. As a
consequence, the maximum total trafc on each link is equal
to the sum of all lower values plus the peak volume of the
largest demand routed on that link. Riis et al. [21] discuss
another application of robust optimization to network design.
D. Robust network design
As anticipated above, the problem at hand asks to dimension
the capacity at each node of G. Each node has to be equipped
with a set of ports, corresponding to interfaces of a network
node with the incident links. A port is capable of routing trafc
to and from that node, and its capacity denes the maximum
trafc in either direction. For example, if a 10 Gbps port is
installed on node i as an interface for link ij, that port will
be able to handle at most 10 Gbps trafc from i to j and 10
Gbps in the opposite direction. Hence, for a trafc of 10 Gbps
i j and 7 Gbps j i there will have to be ports on both
i and j capable of 10 Gbps.
When installing capacity on OTN switches, each demand
is associated with a port and shares the port with no other
demand (see Section II). Hence, we do not need to take into
account demand uncertainty: each demand can be assumed to
be at its maximum value, and therefore trafc uncertainty has
little or no impact on OTN-only network design.
On MPLS links, however, statistical multiplexing substan-
tially reduces the amount of capacity to install. Because more
than one trafc demand is routed on a port, the amount of
capacity to be allocated on each link strongly depends on the
variability of the demand and on the uncertainty model. Hence,
MPLS links are best suited to route uncertain demands as long
as the demands dynamic values can be exploited to reduce
the total trafc w.r.t. the total trafc routed on OTN links, on
which demands are assumed to be xed at their peak value.
We use the uncertainty model of Bertsimas and Sim [19] to
describe the set of trafc matrices. For every q Q we are
given an average demand volume v
avg
q
and a peak volume,
v
peak
q
. Every link (i, j) is used by a set Q
ij
of demands,
of which at most K
ij
can be at their peak value while the
remaining are at average value.
The parameters v
avg
q
, v
peak
q
, and K
ij
are input to our problem
and can be estimated by the network planner based on network
measurements. The accurate estimate of the trafc demands
variability, which is crucial to a correct outcome of the network
planning process, is difcult and has been the subject of
extensive researchsee e.g. the Rocketfuel project [22]. We
assume these parameters to be an input to our problem, and
are likely to be the result of another, equally complex (if not
more complex) analysis task.
The choice of uncertainty model is critical in the design
problem we describe. We have chosen Bertsimas and Sims
model because we make no assumption regarding the trafc
matrix in our experiments, and because one parameter, K
ij
,
allows us to simplify the comparison of different technologies.
The optimization model we propose, nevertheless, admits
a straightforward extension to other uncertainty models as
general as those explained by Ben Ameur and Kerivin [16].
The capacity allocation on all nodes is determined by the
total trafc on the incident links. Consider the set Q
ij
of
demands routed on link ij. Denote as

Q
ij
the subset of the
largest K
ij
demands in Q
ij
, or Q
ij
itself if |Q
ij
| K
ij
. In
order to route all trafc matrices in S, the capacity on ij must
be at least the maximum trafc allowed by S on ij, or

q

Qij
v
peak
q
+

qQij\

Qij
v
avg
q
,
which is the worst-case total trafc among all subsets of trafc
demands routed on ij.
For simplicity, let us assume that uncertainty affects all links
equally, and use a single value K for all K
ij
. Throughout the
paper and in the experiments, we use relatively small values
of K ranging from 1 to 32. We provide in Section III-B an
explanation as to why these small values can tackle realistic
uncertain demands.
E. Sample Network and Trafc
In order to assess the validity and practicality of our
approach, we have created a sample network that closely re-
sembles, in size and demand distribution, a modern backbone
network. The nodes of this network are subdivided into core
nodes and edge nodes. There are 127 nodes in total: 20 core
nodes and 107 edge nodes. Each edge node is connected to
one core node (called its home node) through an edge link,
while each core node is incident to one or more core links.
Figure 1 shows a small sample network with some core and
edge nodes and links.
For all trafc demands, both source and destination are edge
nodes. The routing path of each trafc demand will therefore
contain: one edge link from the source to its home core node;
zero or more core links (zero if the source and the destination
nodes have the same home core node), and an edge link to
the destination node.
We focus on the problem of allocating capacity on core
nodes only, and ignore the capacity to be installed on edge
links and nodes. As will follow from the discussion over
technologies below, the problem of allocating capacity on
edge links admits only one solution and hence is trivial.
Consequently, demands whose source and destination nodes
have the same home core node are ignored in this work.
4
Edge node
Edge link
Core node
Core link
Fig. 1. A two-level network topology.
There are 39 core links, hence there are 146 links in total
(the network has 107 additional edge links, connecting each
edge node with its home core node). Normally, optimization
models on such a large network can be prohibitive, but, since
we can ignore capacity allocation on edge links, we are solving
a problem on a topology of 20 nodes and 39 edges.
The average trafc demands v
avg
q
have been selected by
combining real trafc data from mid-size networks, hence it
represents well the type of trafc that is routed on a network
of this type. There are 4451 trafc demands. The peak value is
assumed, only for our experiments, to be v
peak
q
= v
avg
q
, where
is a constant. Figure 2 shows a distribution of the average
trafc volume of the matrix we have used.
The trafc characteristics used in the model were inu-
enced from different trafc proles from various large service
provider (SP) networks. The traditional SPs have a top heavy
core where 10% of the largest demands occupies about 85% of
the network capacity. On the other hand, cable operators tend
to have atter characteristics with 20% of the largest demands
occupying about 40% of the network trafc. For the purposes
of our model we have the following trafc proles (which
were top heavy but at the same time applicable for different
service providers):
top 10% of demands holds 58% of total trafc;
bottom 40% of demands holds 8% of total trafc;
the total demand is 8 Tbps.
It is worth noting that, although we are interested in the
capacity allocation on core nodes and links, we still consider a
set of 4451 demands and do not group them in macro-demands
dened by a source core node and a destination core node.
This would signicantly reduce the size of the problem, but
would also restrict the set of feasible solutions and therefore
potentially increase the network cost. In fact, consider two
demands q

and q

of Q whose sources and destinations differ


but have the same corresponding source and destination core
nodes: by considering a grouped trafc matrix, all feasible
paths for demands q

and q

would have to be routed as


one single demand and not split on multiple ports, thereby
introducing an implicit and unintended constraint given that
each demand may be routed on either the OTN or the MPLS
layer or in different (smaller) OTN ports.
0
2
3
4
5
6
10
10
10
10
10
1000 2000 3000 4000 5000
Demand
Volume [Mbps]
Fig. 2. The demand volumes (in logarithmic scale) of the 4451 origin-
destination pairs used in our experiments. Note that a large portion of the
trafc is concentrated on a small percentage of the trafc demands.
F. Outline of the paper
Our robust optimization model allocates capacity in a net-
work that uses MPLS and OTN nodes; the network can use
these nodes both standalone and combined. The key question
here is the optimal technology that yields a minimum cost
telecommunication network, given that certain technologies
can take advantage of statistical multiplexing. By applying an
exact solver to the optimization problem, we can get precise
estimates on the cost effectiveness of each technology when
only a limited knowledge of the trafc demand is available.
Our optimization model, while borrowing from other un-
certainty descriptions, is unique in that it addresses a specic
network provisioning problem, where the routing of each
demand is xed in advance but the layer (MPLS or OTN)
where each demand is routed has to be decided.
In the next section, we describe the problem of allocating
capacity on networks that have either MPLS or OTN nodes,
but not both, under the uncertainty model described earlier.
Section III introduces the problem of designing a network
by using both OTN and MPLS nodes and describes a robust
optimization model. In Section IV, we present extensive com-
putational results. We provide conclusions and open questions
in Section V.
II. NETWORK PROVISIONING
We consider four types of capacity, and model how these
technologies result in different cost networks given the as-
sumptions in Section I-A. We then determine the cost-
optimized network based on a combination of the technologies.
Figure 3 shows the structure of the nodes considered: OTN-
only, MPLS-only, combined MPLS+OTN, and bypass.
A. OTN Switching
Optical transport network (OTN) switching technology is
the next generation SONET/SDH (synchronous optical net-
work / synchronous digital hierarchy) technology for providing
5
point-to-point connectivity in a transport network. The OTN
architecture is specied in ITU-T Rec. G.872 and the frame
formats and payload mappings are specied in G.709. OTN is
a circuit switched technology ideal for managing end-to-end
circuit-switched services. The benets of the technology are
that dedicated end-to-end bandwidth is allocated per circuit
offering a single level (but reliable) quality of service (QoS).
As specied in the G.709 specications, individual client
demands are mapped to outer and inner containers called
Optical Data Unit (ODU) and Optical Transport Unit (OTU)
layers for transmission on the transport network. ODU frame
formats depend on the capacity of the physical link (for
example an ODU2 frame is used for 10 Gbps links). Within an
ODUk frame there can be multiple OTUk frames to provide
end-to-end circuit connectivity for different sized circuits.
These can be a combination of OTU0 (1.25 Gbps), OTU1
(2.5 Gbps), OTU2 (10 Gbps), OTU3 (40 Gbps) or OTU4 (100
Gbps). Thus two ow demands of 1 Gbps and 2 Gbps would
map to an OTU0 and an OTU1 frame respectively and then
be wrapped into an ODU2 frame for transmission on the link.
However, in the presence of variation and bursts of ow
demands, circuit-switched networks have to conservatively
allocate bandwidth to handle peak capacity of the ow. This
leads to lower utilization of the circuit and physical link
capacity.
B. Compensating potential aggregation issues with OTN
Smaller-sized ows (less than 1.25 Gbps) are mapped to
ODU0 containers for transport resulting in an acute underuti-
lization of capacity. In the presence of a large number of small
ows, this would lead to a severely inefcient network due
to a large amount of capacity being unused. Filling ODU0s
efciently can only be achieved using an external aggregation
device (adding cost) or via proprietary sub-ODU0 mapping.
The model assumes that such a capability exists in the network,
although this would have to be provided separately.
Another consideration with OTN implementations is how
to reduce the risk of many larger containers (for instance,
ODU1 or ODU2) remaining empty. There is a mechanism
called ODUFlex that some implementations may provide to
concatenate ODUs to create NODU0 up to 100G. The model
accounts for this possibility as well.
C. MPLS routers
Multi-Protocol Label Switching (MPLS) is a packet-
switched technology used for transporting any type of trafc
by grouping it into suitable sized packets. Service levels and
priorities can be set for different types of trafc such as wire-
less, wireline, video and data trafc. A connection oriented
path in a packet-switched network can be set up by means of a
Label Switched Path (LSP) which can be variable-sized based
on trafc needs; thus it is possible to achieve ne-grained
as well as coarse-grained switching of trafc. The existence
of buffers and queuing allows for prioritization of delay and
time sensitive trafc during bursts and statistical variances in
the ow demands. The multiplexing gains thus achieved by
integrating multiple statistically varying demands on a link
provide an efcient utilization of the link bandwidth. Packet-
switched networks are thus most advantageous in networks
where peak demands include multiple higher-than-average
demands.
The functionality we are talking about in Label Switching
Routers (LSRs) is bare label switching and is therefore roughly
equivalent to MPLS Transport Prole (MPLS-TP). These
ports are usually more expensive than OTN ports, therefore
a tradeoff arises between relatively expensive equipment that
can exploit statistical multiplexing and cheaper equipment,
which is unable to take advantage of statistical multiplexing
and therefore has to be installed in larger quantities. Hence
we are interested to know, for given uncertainty sets and an
initial network topology, which technology among MPLS or
OTN is more cost effective.
D. Bypass
In a typical network of MPLS routers, trafc enters the
network through the ingress router and transits over zero or
more routers in reaching the egress router. In a fully meshed
network, all nodes are connected to each other and thus
trafc from an ingress router directly travels to the egress
router. However, in a typical network, as the number of nodes
increases it is quite complex to connect each node to every
other node; as a result trafc transits over some routers en
route from the ingress to the egress node. Each node thus has
a portion of trafc that is local (i.e. is sourced or destined to
that node) and the rest of the trafc is transit trafc (destined
to another node). Since router ports are needed for both local
and transit trafc, one can reduce the number of router ports
for transit trafc by introducing an OTN layer between the
routers. The MPLS routers switch only the local trafc; transit
trafc at a node bypasses the router and is switched by the
OTN layer. Since the OTN layer aggregates local and transit
trafc, additional trunking ports are needed between the OTN
and MPLS switches.
E. Provisioning at the MPLS, OTN, and Bypass layers
Before introducing the optimization model used in our
experiments, we show how to allocate capacity on MPLS-
only, OTN-only, or bypass networks. This is straightforward
and does not need any optimization process. In the following,
for each trafc demand q Q we dene the path p
q
as the set
of directed links on which q is routed, determined by solving
a shortest path problem on the OSPF weights.
We rst show how to compute the capacity of an OTN-
only network. We assume that 1.25 Gbps, 2.5 Gbps, and 10
Gbps ports can be installed at the OTN layer. As mentioned
in Section II-A, 10 Gbps links have to be installed to accom-
modate the set of ports. Each demand q Q, whose volume
must be assumed to be xed at v
peak
q
, is routed on a port with
an equal or larger capacity. Therefore, the number of ports
can be determined immediately. Consider the vector of OTN
port types, c
otn
= (1.25, 2.5, 10). The number of OTN ports of
capacity c
otn
k
, with k {1, 2, 3}, to be installed at the interface
of node i to node j in V is the maximum between the ports
6
(a) OTN switch (b) MPLS router (c) Combined MPLS+OTN (d) Bypass router
Fig. 3. The structure of OTN, MPLS, combined OTN and MPLS, and bypass nodes.
used for incoming trafc and those for outgoing trafc: for
each k {1, 2, 3},
n
otn
ij,k
= max{|{q Q : (i, j) p
q
c
otn
k1
< v
peak
q
c
otn
k
}|,
|{q Q : (j, i) p
q
c
otn
k1
< v
peak
q
c
otn
k
}|},
where c
otn
0
= 0. In order to nd the amount of 10 Gbps
OTN links to be allocated, note that in our problem denition
the port capacities are multiples of one another, and are sub-
multiples of the 10 Gbps link capacity. This simplies the
computation of the number N
otn
ij
of links:
N
otn
ij
=
1
10
max{1.25n
otn
ij,1.25
+ 2.5n
otn
ij,2.5
+ 10n
otn
ij,10
,
1.25n
otn
ji,1.25
+ 2.5n
otn
ji,2.5
+ 10n
otn
ji,10
}.
It is then straightforward to compute the total number of 10
Gbps links at each node i V , which is given as the sum of
N
otn
ij
over all j adjacent to i.
For allocation of small demands into ODU0, the number
of ports of 1.25 Gbps is determined as follows. Consider all
demands whose volume v
peak
q
is below 1.25 Gbps, and routed
on 1.25 Gbps ports on link ij in the direction from i to j,
and denote it as Q
1.25
ij
. A set of ports have to be allocated
so as to accommodate all demands of Q
1.25
ij
, in such a way
that multiple demands of Q
1.25
ij
can be routed on the same
port, although they must be entire demands. Note that this
a signicant change with respect to the OTN ports, each of
which is instead dedicated to one demand.
Therefore, one has the problem of minimizing the number
of 1.25 Gbps ports that can accommodate some or all demands
below 1.25 Gbps. This is equivalent to solving a bin packing
problem [23] on every link. More specically, Q
1.25
ij
has to
be partitioned into H subsets Q
1.25
ij,h
, for h 1, 2 . . . , H, such
that

H
h=1
Q
1.25
ij,h
= Q
1.25
ij
and

qQ
1.25
ij,h
v
peak
q
1.25 Gbps for
each h = 1, 2 . . . , H. Given that we need H such ports, a small
value of H is desirable to keep the network cost limited. The
design process will thus require the solution of one bin packing
problem for each link ij. We do not provide further details for
the sake of conciseness, but point out that these problems are
relatively easy for the instance size we are considering.
Once this problem is solved, the number of ports for each
link ij on both directions i j and j i is known and
denoted as n
odu
ij
and n
odu
ji
respectively. The number of ODU0
ports to be installed on node i is given by

jV :(i,j)E
max{n
odu
ij
, n
odu
ji
}.
In the MPLS case, the capacity depends on the average
and peak values of the demands. We assume that only 10
Gbps ports are available at the MPLS level, and compute the
number of such ports based on the set of demands routed
on link ij in the direction i j, denoted Q
ij
, and in the
opposite direction, denoted Q
ji
. We recall that these subsets
of demands are known in advance. The total MPLS trafc on
a link ij under our uncertainty scenario has been described in
Section I-D, and we report the number of 10 Gbps ports:
n
mpls
ij
=
_
1
10
_

q

Qij
v
peak
q
+

qQij\

Qij
v
avg
q
__
.
The bypass case is slightly more complex: the incoming
and outgoing trafc at each node i are sent onto the network
through an MPLS router that allows statistical multiplexing.
That trafc is then forwarded on the network by means of OTN
switches. At each node, transiting trafc is handled by OTN
switches only, while incoming and outgoing trafc traverses
MPLS routers. Suppose Q
src
i
is the set of demands with origin
i, and Q
dst
i
the set of demands with destination i. Denote as

Q
src
i
the set of K largest demands among

Q
src
i
, or Q
src
i
itself
if |Q
src
i
| K, and analogously dene

Q
dst
i
. The total capacity
7
to be installed is the number of ports for the local trafc
n
byp,local
i
=
_
1
10
max
_
q

Q
src
i
v
peak
q
+

qQ
src
i
\

Q
src
i
v
avg
q
,

q

Q
dst
i
v
peak
q
+

qQ
dst
i
\

Q
dst
i
v
avg
q
__
,
which has to be considered both at the MPLS layer and at the
OTN layer (i.e., the cost of each such port will be that of an
OTN port plus an MPLS port) plus the number of OTN ports
for the transit trafc, n
byp,tr
ij
. This, in turn, is the total, among
all adjacent nodes j, of the ports to be installed at node i
for trafc owing on link ij in both directions. Let us denote
Q
tr
ij
the set of trafc demands routed on ij from i to j not
originating in i, and

Q
tr
ij
Q
tr
ij
a subset containing the K
largest demands (or, again, Q
tr
ij
itself if |Q
tr
ij
| K). Denote
Q
tr
ji
the set of trafc demands routed on ji from j to i not
ending in i, and dene

Q
tr
ji
analogously. Then
n
byp,tr
ij
=
_
1
10
max
_
q

Q
tr
ij
v
peak
q
+

qQ
tr
ij
\

Q
tr
ij
v
avg
q
,

q

Q
tr
ji
v
peak
q
+

qQ
tr
ji
\

Q
tr
ji
v
avg
q
__
.
In summary, in the OTN-only, bypass, and the MPLS-only
scenarios the capacity allocation can be carried out by a simple
procedure which does not require solving an optimization
problem. Relatively small optimization problems have to be
solved for the allocation of demands to the single OTU0 links.
One such problem must be solved per link in order to compute
the total capacity on that link.
III. OPTIMIZATION MODELS FOR COMBINED MPLS-OTN
SOLUTIONS
In order to cut costs even more, one could think of building
a network with both MPLS and OTN ports that can route
all trafc matrices in the uncertainty set. Hence, the network
admits both MPLS routers and OTN switches at the same
hierarchy, and the problem is that of determining the number
of MPLS and OTN ports for all nodes of the network.
Because we are dropping the restriction to use only one
technology, the resulting network, excluding additional costs
to interface, if necessary, MPLS routers to OTN switches, will
be less expensive than both the OTN-only and the MPLS-
only network. The question is then how much cheaper this
combined network would be.
This network optimization problem can be shown to be NP-
hard, i.e., it is very unlikely that there be an algorithm that
solves it in a number of elementary steps that is polynomial
in the size of the problem (number of nodes and links of G).
The proof of NP-hardness is by reduction from the Subset
Sum problem [24]: consider a set S and a function c : S R.
Dene c(S

) =

iS
c(i) for any S

S. The Subset Sum


problem asks to partition S into two sets S
1
and S
2
(i.e.,
S
1
S
2
= S and S
1
S
2
= ) in such a way that |c(S
1
)c(S
2
)|
is minimized. Our network design problem is NP-hard because
any instance of Subset Sum problem on a set S and a function
c can be transformed in an instance of our problem, where the
network G has only one link, the number of trafc demands
equals |S|, = 1, K = 0, and the trafc value of each demand
i S is c
i
. We omit the details for the sake of conciseness.
This general design problem can be formulated as a Mixed
Integer Linear Programming (MILP) model [25]. The MILP
model illustrated in this section generalizes the design method
outlined above for OTN-only and MPLS-only networks: by
forbidding either the MPLS or OTN variables in the model
below, we obtain a much simpler problem that admits the
solution outlined above.
A large portion of demands have routing path spanning
two or more core links. These demands, in principle, could
use the MPLS layer on certain links of the routing path and
the OTN layer on the remaining links. This would imply
that some intra-node, inter-layer capacity is needed on one
or more nodes. However, we have strong empirical evidence
suggesting that solutions where a demand uses both layers
is sub-optimal. More specically, preliminary experiments on
the aforementioned 20-node network, for multiple values of K
and , have shown that there is at least one optimal solution
where each demand is routed on one layer. A consequence of
this is a major simplication of the optimization model and
thus shorter solution times. Therefore, from now on we assume
that each trafc demand is entirely routed on either the MPLS
layer or the OTN layer and a key decision to make is which
layer each demand is routed on.
Suppose that Q
mpls
ij
is the set of demands routed on the
MPLS layer of link ij, and analogously Q
otn
ij
for the OTN
layer. In this case, the amount of ports to be allocated for i and
j is given in Section II-E. Because this quantity depends on
Q
mpls
ij
and Q
otn
ij
, which are not known in advance, the decision
on each demand is a variable in an optimization model.
For each trafc demand q Q, we dene binary variables
f
otn
q
, f
odu
qh
, and f
mpls
q
. Variables f
otn
q
and f
mpls
q
are one if demand
q is routed on the OTN or MPLS layer, respectively, and
zero otherwise. Variable f
odu
qh
is one if demand q uses the
subset Q
1.25
ij,h
(see denition in the previous section), and zero
otherwise. Here the number H of 1.25 Gbps ODU ports is
computed so that enough capacity is available in case all of
the demands in Q
1.25
ij
are routed on the ODU layer, but we
can expect that not all of the variables f
odu
qh
are nonzero. These
binary variables are subject to the following constraint:
f
otn
q
+f
mpls
q
+

hH
f
odu
qh
= 1 q Q, (1)
which forces exactly one of these variables to one and the
remaining to zero. There are |Q|(H + 2) such variables.
We also dene a set of integer variables for the ports on
the OTN layer: x
otn
ij,t
is the number of OTN ports on link ij
(i.e., installed at i and j), where t indexes the capacity vector
c
otn
= (1.25, 2.5, 10). Also, we dene binary variables x
odu
ij,h
,
which are one if 1.25 Gbps ODU0 ports are required to support
the combination of ports in h for aggregated demands. Finally,
the integer variable x
mpls
ij
is the number of 10 Gbps MPLS ports
allocated on link ij.
These variables are related to ow variables through the
capacity constraints. At the OTN layer, we need one OTN port
of capacity 1.25 Gbps on link ij for each demand q routed
on ij whose peak value is at most 1.25 Gbps and such that
f
otn
q
= 1. Similarly, we need one 2.5 Gbps OTN port on link
ij for each demand q routed on ij whose peak value is strictly
greater than 1.25 Gbps and not larger than 2.5 Gbps, and such
that f
otn
q
= 1, and similar considerations apply to 10 Gbps
8
OTN ports. We model this with the constraint
x
otn
ij,t

q Q : ij pq
c
otn
t1
< vq c
otn
t
f
otn
q
ij E, t {1, 2, 3}, (2)
where c
otn
0
= 0. The capacity constraint for ODU ports requires
that a port be present if there is nonzero trafc on it:
1.25 x
odu
ij,h

qQ:ijpq
v
peak
q
f
odu
qh
h = 1, 2 . . . , H, ij E.
(3)
Finally, the capacity on each link of the MPLS layer is dened
as follows:
10 x
mpls
ij

qQ
v
avg
q
f
mpls
q
+z
ij
ij E, (4)
where z
ij
is a variable describing the peak ow, and is
dened by a special port capacity constraint. Let us dene
the peak-average demand difference as
q
= v
peak
q
v
avg
q
.
Then the peak ow is given by the optimal solution value of
another optimization model which yields the worst-case (i.e.,
maximum) trafc on ij as a function of the set of demands
routed at the MPLS layer of ij:
ij E z
ij
max

q

Q:ijpq

q
f
mpls
q
s.t.

Q Q
|

Q| K.
(5)
Given that only a subset of demands of cardinality K is al-
lowed to have peak value, the right-hand side of this constraint
can be rewritten as
max

qQ:ijpq

q
f
mpls
q

q
s.t.

qQ

q
K

q
{0, 1}q Q,
which is an optimization model on binary variables
q
dening
the set of demands at peak value. In order to evaluate constraint
(5) one has to solve, for each link ij, a maximization problem
in variables
q
where f
mpls
q
are taken as parameters.
Another class of variables represents the number of links
needed at all nodes, which constitutes the ultimate decision
variable on which the total network cost is computed. We
dene the variables y
otn
ij
as the number of 10 Gbps links to
be installed on both ends of link ij at the OTN layer. The
constraints relating link variables to port variables are easily
dened, for each ij E, as follows (note that OTN links must
accommodate OTN and ODU ports):

t{1,2,3}
c
otn
t
x
otn
ij,t
+ 1.25
H

h=1
x
odu
ij,h
10 y
otn
ij
ij E. (6)
Each 10 Gbps OTN link must be assigned a set of whole
OTN ports, i.e. trafc on an OTN port cannot be shared
between two OTN links. In general, a more complicated set
of constraints, resembling the bin packing set of constraints
for ODU port/ow capacity, would be necessary. However, the
link capacity, 10 Gbps, is a multiple of the largest OTN port
capacity, and each OTN port capacity c
otn
t
is a multiple of the
preceding OTN port capacity, c
otn
t1
, thus resulting in another
bin packing problem that is implicitly solved in (6).
This set of variables and constraints is only dened at the
OTN layer, as the number of MPLS links is equal to that of
MPLS ports x
mpls
ij
, hence no additional variable is needed.
Let us assume that the cost of a single link is g
otn
at the OTN
layer and g
mpls
at the MPLS layer. It is realistic to assume that
g
mpls
= g
otn
with 1, i.e. the cost of MPLS links is greater
than that of OTN links. We have set = 1.3 in most of our
experiments, which reects a cost differential that seems more
realistic for the kind of functionality required, at the MPLS
level, from this type of network. This introduces a trade off
between the cheaper (but under-utilized) OTN links and the
more expensive MPLS links, which use statistical multiplexing
and can hence make a better use of the total capacity. The
objective function is the total core network cost, i.e., the total
OTN link cost plus the total MPLS link cost
2

ijE
(g
otn
y
otn
ij
+g
mpls
x
mpls
ij
), (7)
where all quantities are multiplied by two because the number
of ports on link ij is installed at both end nodes i and j.
To recap, the problem P of determining the minimum cost
capacity on the nodes of G to route all demands in the
uncertainty set S requires to minimize the objective function
(7) subject to constraints (1,2,3,4,5,6). This is a nonlinear
integer optimization problem, the only nonlinear constraint
being (5).
A. Robust optimization: an opponents view
Problems like P are in general classied as bilevel program-
ming problems as the evaluation of a constraint is associated
with solving another optimization problem [26]. Our model is
representative of a smaller class of bilevel programming, that
of robust optimization. The key component is the right-hand
side of constraint (5), which, as said before, is evaluated by
solving a linear optimization problem. The interpretation of
this inner optimization problem is functional to understanding
the advantages and the meaning of robust optimization.
Let us suppose an external player (such as nature, competi-
tors, or customers) is given the solution to our optimization
model (1,2,3,4,5,6,7) (i.e., the decisions on the layer on which
every demand is routed) and has the intent of maximizing
the nuisance on the network we have designed. This external
player, which is in practice our opponent, has the control over
the uncertainty set, i.e., decides which demands in a subset of
cardinality K are at their peak value. The opponent will then
determine a subset of demands that maximizes the load on the
installed capacity. In order to anticipate the opponents move,
we must implicitly solve the optimization model.
We begin by relaxing integrality over variables
q
and
constrain them in the interval [0, 1] rather than in the non-
convex set {0, 1}. This is a relaxation of the above problem
in that all solutions that are feasible for the initial problem
are also feasible when requiring that
q
be continuous. Also,
because the coefcient matrix of the problem (5) is totally
unimodular, any optimal solution of the relaxed problem is
optimal for the original problem as well [25].
Now the maximization problem on the right-hand side of
constraint (5) is a Linear Programming (LP) problem, which
9
is much easier to handle for our purposes. We rewrite it here
for the sake of clarity:
P
ij
: max

qQ:ijpq

q
f
q

q
s.t.

qQ

q
K
0
q
1 q Q.
In order to get rid of the maximization sign, we use weak
duality and apply a well known procedure in robust optimiza-
tion [27]. Any LP problem max{c

x : Ax b}, with an
n-vector of variables x R
n
and m constraints, with c Q
n
,
A Q
mn
, and b Q
m
, admits a dual LP problem
min{u

b : u

A = b, u 0}
with a vector u R
m
of variables. Weak duality dictates that
for any vector x satisfying Ax b and for any non-negative
vector u satisfying u

A = b, we have [28]
u

b c

x.
As a consequence, u

b provides an upper bound to the right


hand side of constraint (5). Replacing the right-hand side of
(5) with u

b and adding the dual constraints u

A = b, u
0 to the original problem allows us to implicitly solve the
maximization problem in (5) and obtain an upper bound on the
total trafc on link ij. It also yields a set of linear constraints,
thus eliminating the nonlinearity introduced by the max sign.
Consider each link ij E and the associated problem P
ij
.
Denote
ij
the dual variable of constraint

qQ

q
K and

q
ij
the dual variable of the upper bounding constraint for
q
.
Then the dual of problem P
ij
is the minimization problem
D
ij
: min K
ij
+

qQ

q
ij
s.t.
ij
+
q
ij

q
f
q
ij E, q Q

ij
0 ij E

q
ij
0 ij E, q Q.
This yields a model where constraints (5) are replaced by the
following sets of constraints:
z
ij
K
ij
+

qQ

q
ij
ij E

ij
+
q
ij

q
f
q
ij E, q Q

ij
0 ij E

q
ij
0 ij E, q Q.
Thus, we replace the set of |E| nonlinear constraints by
introducing |Q||E| +|E| new variables and |E| +|Q||E| con-
straints. Although this increases the model size substantially,
eliminating the nonlinear constraint (5) reduces the problem
to an Integer Linear Programming (ILP) problem, thus more
tractable. Also, it can be shown that only a small subset of
these constraints are necessary, by replacing Q with Q
ij
in
the above models P
ij
and D
ij
. We omit the details for the
sake of conciseness, but we did implement this reduction in
our experiments.
As an interpretation of this procedure, consider the initial
problem and remove the max term in constraint (5), and
suppose that
q
instead are variables of P. This would mean
that we, not the opponent, control
q
. Then we no longer
have to explicitly solve an optimization model, and can decide
which demands assume peak value. Solving such a (nonlinear)
problem yields an unrealistic best-case solution, in which the
most optimistic set of demands assume peak value: the empty
set. In fact, it is easy to prove that any optimal solution of such
a problem would have all
q
set to zero, therefore obtaining
a null value of z
ij
, which corresponds to a no-peak scenario.
Because we use the dual of this problem, by eliminating
the minimization sign we still impose the dual constraints,
thus forcing the variables
q
ij
and
ij
to be dual-feasible. This
in turn ensures that, for any feasible solution of the overall
problem, the right-hand side of constraint (5) is an estimate
from above of the maximum capacity allocation, hence forcing
z
ij
to a value that is no smaller than the optimal solution of
problem P
ij
. This yields a safe estimate (from above) of the
total peak trafc on link ij. Eliminating the minimization sign
puts the dual variables
q
ij
and
ij
in our control, albeit forcing
them to model a worst-case scenario.
B. Statistical multiplexing and robust allocation
The fundamental assumption of the model outlined above
is that the uncertainty in the trafc demand can be dealt with
by allocating capacity to accommodate a set of demands at
their average value plus a worst-case extra capacity dened by
the largest K demands. Given that this worst-case total trafc
is computed independently at every link, the total capacity
allocated on the network may be overly conservative even
for small values of K. This suggests that an appropriate
value of K is not easy to nd. As discussed above, the
availability of statistical data for trafc matrices may provide
more information, though it would be difcult to incorporate
it in an optimization model as simple as the one described.
Let us consider, in the remainder of this section, a different
problem: how does a value of K compare to a random
set of demands at their peak values? In other words, for a
capacity allocation that includes extra ports for supporting the
largest K peaks, can we estimate the number T of demands
that, on average, will be satised by the resulting capacity
allocation? Note the contrast between K and T: while the
former determines the set of largest demands routed on a
link, the latter is the expected number of demands, randomly
selected from Q, that can have peak value and still are routed
on every link.
Suppose that each demand can take its peak value with
equal probability, that is, each demand has probability T/|Q|
of being at peak value. The total extra capacity can be
computed as the expected value of the extra demand using
the uniform distribution p
q
= T/|Q| for all q Q, and is
D =

qQ

q
T
|Q|
. In order to compare D with the worst-case
amount of extra capacity to install, we must have

qQ

q
T
|Q|
=

q

Q

q
,
where

Q is the set of K largest demands. An estimate of T
is therefore
T =
|Q|

q

Q

qQ

q
.
Suppose now that the probability that a demand is at peak
value grows with the demand volume, and let us denote this
10
probability distribution as progressive. In particular, assume
w.l.o.g. that v
avg
1
v
avg
2
... v
avg
|Q|
. Suppose, furthermore,
that the probability p
q
that a demand q Q be at its peak
value is proportional to its index q, i.e. p
q
= q, where is
such that the expected number of demands at peak values is
T, i.e.

qQ
q = T and hence
=
T

qQ
q
=
2T
|Q|(|Q| + 1)
.
Again, we obtain T as a function of K by solving the equation

qQ
p
q

q
=

q

Q

q
,
which yields
T =
|Q|(|Q| + 1)

q

Q

q
2

qQ

q
.
The following table provides values of T for small values of
K for both the uniform and the progressive case using demand
data from the instance we have used.
K 1 2 4 8 16 32
T (unif.) 160 210 300 446 642 895
T (prog.) 96 126 180 268 385 587
When allocating capacity, the worst-case scenario, even for
small values of K, can cover a number of demands at their
peak value on a single link ij that is much larger than K, and
that can be up to one fth of |Q|, for K = 32 in the uniform
case. This table would clearly look different if the demand
distribution were different: the instance we have chosen has
a number of very large demands, while the volume is slowly
degrading for the smaller ones.
In practice, the number of simultaneous peaks in a real
network would likely be up to 10. However, for the purpose
of meaningful analysis, we have considered K up to 32 peaks.
Note that a larger K means that more bandwidth needs to be
provisioned in an MPLS network and thus a higher cost.
IV. COMPUTATIONAL RESULTS
We present below the results of a set of experiments
conducted on the robust optimization models described in this
paper. The main purposes of these experiments are:
1) to assess the utility of the proposed robust optimization
models as a means to obtain networks that are resilient
to at most K simultaneous peaks;
2) to compare networking technologies in terms of total
network costs; and
3) to quantify the dependence of the network cost on the
number K of simultaneous peaks allowed;
K takes values in the set {1, 2, 4, 8, 16, 32}. These values are
only apparently small in comparison with |Q| = 4451: apart
from the reasons pointed out in the previous section, we also
have observed, in preliminary tests, that a further increase of
K does not change the results signicantly: in several cases,
increasing K from 32 to 64 yields an increase of 1% or less
on the network cost, while changing K from 1 to 2 leads to
a steeper increase. Furthermore, numerous other experiments
have been conducted previously with smaller values of K, and
in practice even 32 is considered very conservative.
As we assume that the ratio between peak and average
volume is a constant 1, i.e., v
peak
q
= v
avg
q
q Q,
all trafc matrices are described by the set
S = {v
avg
q
(1 + ( 1)
q
), q Q :

qQ

q
K,
q
{0, 1}q Q}.
This denition of uncertainty is uniquely determined by
and K. Wrapping the denition of uncertainty around just
two parameters may seem too simplistic, but this allows us
to formulate a compact optimization model for the capacity
allocation problem. In order to accurately describe the real
uncertainty set, appropriate values of K and are necessary.
It is barely worth noting here that the larger K and , the more
conservative the result, i.e., the more expensive the network:
for very large K, we are allowing many trafc demands to
be at their peak value, therefore increasing the potential total
trafc on each link. If K = |Q|, the uncertainty set can
actually be reduced to a single point, as S is dominated by
a single trafc matrix: one in which all demands are at their
peak value (for a detailed description of dominance in trafc
demands, see [29]). The opposite case is one in which K = 0
(or equivalently = 1), where S is given by the trafc matrix
in which all trafc demands are equal to their average value.
A. Implementation details
We have implemented our optimization models using the
AMPL modeling language [30], and solved them with the
Gurobi Mixed Integer Linear Programming solver
1
, a state-of-
the-art commercial solver which implements a parallel branch-
and-bound algorithm. All experiments were conducted on a
Linux 64 bit machine with four processors, which are used
in parallel by the branch-and-bound algorithm, i.e., when the
branch-and-bound solver is operative, at most four branch-and-
bound nodes are solved simultaneously. All solver parameters
were set to their default value.
Integer Linear Programming problems are difcult, and
no polynomial-time solver is known for them [25]. As a
consequence, if the time allowed is limited, ILP solvers can
solve to optimality i.e., provide a solution with cost z
opt
and
a proof that it is optimal only instances of a certain size.
If the time allotted to search for an optimal solution is
scarce, ILP solvers attempt to provide a feasible, sub-optimal
solution, with cost z
feas
z
opt
, and are always able to provide
a lower bound z
lb
z
opt
of the optimal solution. Depending
on the quality of the solver, the optimality gap
zfeaszlb
zlb
(which
is zero when an optimal solution is found) provides a measure
of the quality of the solution found. In all our experiments, we
have imposed a time limit of two hours to all ILP instances
solved, namely those for the combined MPLS+OTN network.
We have observed that in all of these tests, either an optimal
solution was found or a relatively good one was found with
an optimality gap below 1%.
1
See http://www.gurobi.com for more information.
11
B. Capacity allocation
Table I shows the results for the four technology combi-
nations considered: Bypass, Optimized MPLS+OTN, MPLS
only, and OTN only. For each value of K {1, 2, 4, 8, 16, 32}
and {1.5, 3, 4.5}, we report the network cost (7) for all
four alternatives and the corresponding number of 10 Gbps
links that need to be installed on each network. For the bypass
network, we also report the local trafc and the transit trafc
(columns 3, 4, and 5, respectively). It is evident that MPLS
links, which fully exploit statistical multiplexing, use network
resources more efciently and hence limit the increase in
network cost when the uncertainty increases (i.e. for large
values of K and ).
In Table II we provide similar results with one parameter
change: the cost ratio between OTN and MPLS is set to one,
i.e., OTN and MPLS links are assumed to have equal cost in
this case. It comes as no surprise that the tradeoff between
MPLS routers and OTN switches favors MPLS networks even
more: when MPLS links and OTN links have equal cost, a
network composed solely of MPLS links is much cheaper
than the corresponding OTN network. We omit the results for
the optimized (combined) MPLS+OTN network as they are
unnecessary in this case: the combined network has an optimal
solution equal to that of the MPLS-only network.
Bypass MPLS OTN
Links
K transit local cost cost links cost links
1.5 1 767 1750 32840 17420 1742 38460 3846
2 787 1820 33940 18100 1810
4 816 1898 35300 19040 1904
8 854 1986 36940 19740 1974
16 906 2098 39100 21060 2106
32 958 2218 41340 22060 2206
3 1 846 1996 36880 19900 1990 69340 6934
2 927 2236 40900 22260 2226
4 1048 2556 46520 25560 2556
8 1201 2930 53320 29340 2934
16 1398 3370 61660 33700 3370
32 1614 3836 70640 38300 3830
4.5 1 927 2246 41000 22380 2238 98980 9898
2 1074 2672 48200 26720 2672
4 1282 3232 57960 32320 3232
8 1549 3890 69880 38940 3894
16 1894 4646 84340 46460 4646
32 2269 5482 100200 54640 5464
TABLE II
COMPARISON OF NETWORK COSTS FOR SEVERAL TRAFFIC UNCERTAINTY
SETS, WHERE MPLS AND OTN LINKS HAVE THE SAME COST.
Finally, in Table III we summarize the previous two tables
by pointing out the percentage difference in cost obtained
when using MPLS nodes rather than OTN nodes. Note that
the the cost difference is large when =
gmpls
gotn
= 1, for
higher values of and low values of K (see e.g. the 77.3%
highlighted in the table), and viceversa the difference is
relatively small (although still more than 30%) when the cost
ratio is 1.3 and for small values of and large values of K.
C. Approximating worst-case scenarios using xed peaks
Another perspective on the robust optimization model pre-
sented above arises from considering a single scenario, albeit
OTN MPLS
K = 1 K = 2 K = 4
cost cost % cost % cost %
1.3 1.5 38460 22466 41.5 23170 39.7 23958 37.7
1.3 3 69340 24928 64.0 28784 58.4 32872 52.5
1.3 4.5 98980 29020 70.6 33424 66.2 41608 57.9
1 1.5 38460 17420 54.7 18100 52.9 19040 50.4
1 3 69340 19900 71.3 22260 67.8 25560 63.1
1 4.5 98980 22380 77.3 26720 73.0 32320 67.3
K = 8 K = 16 K = 32
cost cost % cost % cost %
1.3 1.5 38460 24396 36.5 25382 34.0 25980 32.4
1.3 3 69340 37072 46.5 41834 39.6 45984 33.6
1.3 4.5 98980 49340 50.1 58034 41.3 65848 33.4
1 1.5 38460 19740 48.6 21060 45.2 22060 42.6
1 3 69340 29340 57.6 33700 51.3 38300 44.7
1 4.5 98980 38940 60.6 46460 53.0 54640 44.7
TABLE III
COST REDUCTION OF AN MPLS NETWORK WHEN COMPARED TO AN OTN
NETWORK. NOTE THAT =
gMPLS
gOTN
.
a very conservative one: the K largest demands are set to their
peak value, while the remaining ones are at average value.
This is equivalent to restricting the uncertainty set S to
one single trafc matrix, which is very limiting from the
opponents standpoint given that many scenarios are ruled out.
The clear advantage is that robust optimization is not necessary
to model this, as the trafc matrix simply needs to be set
according to the single scenario.
Extensive tests performed for all values of K to compare
the total network cost, in both cases, indicate that assuming
a single, although pessimistic, scenario produces very cheap
networks compared to the more conservative robust optimiza-
tion model. This can be explained by projecting the K largest
demands onto their routing paths, which probably do not cover
the whole set of links E and contribute a very small increase
of capacity on each link. This heuristic method can prove very
effective in larger networks if a solution is needed in shorter
time, but clearly no guarantee is provided on all other trafc
matrices of the uncertainty set. Other realistic scenarios, where
a disjoint set of smaller demands are at their peak value, would
overload the capacity installed on those links that are not used
in the routing paths of the largest K demands.
D. Equal Cost Multi-Path (ECMP) routing
Up to this point, we have assumed that all routing paths
follow the OSPF protocol and are established as shortest paths
according to weights provided as input. In order to obtain
Tables I-III, we have performed tests with single routing paths,
i.e., we did not impose the Equal Cost Multi-Path (ECMP),
which splits part of the trafc if more than one shortest path
to destination is available.
We have also performed some tests (not reported here) by
rst imposing ECMP at the MPLS layer only and then at both
layers, to address the effects on the network cost, and we
have observed no substantial difference. The OSPF weights
provided with the instance were such that few point-to-point
demands admit more than one shortest path. More precisely,
there are 20(20 1) = 380 pairs of core nodes in the core
network. Of these, only 10 pairs admitted more than one
12
Bypass Optimized MPLS OTN
Local trafc (Mbps) Links Links
K in out transit MPLS OTN cost cost OTN MPLS cost Links cost Links
1.5 1 7248.3 7236.1 8595.9 767 1750 35141 22466 86 1662 22646 1742 38460 3846
2 7451.6 7423.2 8893.2 787 1820 36301 23170 458 1430 23530 1810
4 7738.3 7698.4 9289.9 816 1898 37748 23958 828 1206 24752 1904
8 8079.6 8077.0 9754.9 854 1986 39502 24396 994 1112 25662 1974
16 8534.0 8559.4 10295.2 906 2098 41818 25382 1038 1154 27378 2106
32 9051.8 9068.0 10894.2 958 2218 44214 25980 1116 1140 28678 2206
3 1 8037.1 7988.5 9831.1 846 1996 39418 24928 106 1836 25870 1990 69340 6934
2 8850.5 8737.0 11017.0 927 2236 43681 28784 86 2148 28938 2226
4 9997.2 9837.4 12607.5 1048 2556 49664 32872 240 2344 33228 2556
8 11362.5 11351.8 14472.7 1201 2930 56923 37072 868 2184 38142 2934
16 13179.7 13281.1 16637.3 1398 3370 65854 41834 1482 2078 43810 3370
32 15250.6 15315.4 19023.3 1614 3836 75482 45984 2482 1628 49790 3830
4.5 1 8826.0 8740.8 11067.6 927 2246 43781 29020 94 2160 29094 2238 98980 9898
2 10249.5 10050.8 13146.3 1074 2672 51422 33424 134 2468 34736 2672
4 12256.1 11976.4 15964.7 1282 3232 61806 41608 344 2936 42016 3232
8 14645.4 14626.7 19263.0 1549 3890 74527 49340 1190 2880 50622 3894
16 17825.6 18003.1 23053.9 1894 4646 90022 58034 2010 2918 60398 4646
32 21449.7 21563.1 27219.6 2269 5482 107007 65848 3470 2396 71032 5464
TABLE I
COMPARISON OF NETWORK COSTS FOR SEVERAL TRAFFIC UNCERTAINTY SETS. THE MPLS/OTN RATIO OF LINK COST IS 1.3, I.E., MPLS PORTS COST
30% MORE THAN OTN PORTS. NOTE THAT THE DATA FOR OTN NETWORKS DOES NOT DEPEND ON K.
shortest path. Because the experiments under ECMP did not
show big changes and because ECMP is outside the scope of
this work, we omit the full results.
E. Reducing the overall demand upon multiple peaks
The set S of potential trafc matrices, a crucial concept in
our model, is in general a polyhedron and can be modied if
needed. We consider an example of uncertainty set that avoids
over-conservativeness.
The model we have discussed so far assumes that at most
K demands may take on a peak value. At any moment, if
the number of demands at peak value is small compared to
K, link usage (and thus congestion) is well below the limits
determined by the network capacity. However, if close to K
demands are at peak value, congestion and capacity occupation
gures can be critical and network service might be at risk.
To prevent this risk, or to limit the increase in network cost
associated with large K, the network planner and the network
users may enter contractual agreements dictating that all net-
work users accept to reduce their demand, no matter whether
it is at peak value or not, by a given (small) percentage, which
varies with the number of current peak demands. In other
words, if, at a certain instant, few or none of the demands
are at peak value, the network functions normally and users
can communicate at the requested data rate. However, with a
growing number of peak demands, all users will transmit at
a slightly reduced rate. A reduction dened by a parameter
(typically {5%, 10%}) can be considered, and it is fully
enforced when the number of peak demands is exactly K. As
a result, total network throughput can be decreased articially
by reducing the volume of all demands, while, if all demands
are at average value, they are left unchanged.
The rationale behind this change in the uncertainty set S is
the need to reduce the cost of a network whose trafc demands
are, most of the time, all at average value. In the relatively rare
bursty periods, with K demands at peak value, it becomes
reasonable to decrease every demand for the (presumably
small) duration of the burst phase. If K

< K demands are at


their peak value, a smaller reduction of
K

K
can be enforced.
This uncertainty set can be modeled with linear constraints.
Dene a set of binary variables
q
, for each q Q, indicating
whether demand q is at average value (
q
= 0) or peak value
(
q
= 1), and a variable Z equal to the ratio of the number of
demands at peak value to K. The uncertainty set is dened by
the set of vectors v R
|Q|
+
subject to the following constraints
(which will be replicated for all ij E):
Z =

qQ

q
/K (8)
v
q
(1 Z)v
avg
q
q Q (9)
v
q
(1 Z)v
avg
q
q Q (10)
v
q
(1 + ( 1)
q
)v
avg
q
q Q (11)

qQ

q
K (12)

q
{0, 1} q Q.
These constraints reduce to the simpler set of constraints of
the maximization model in (5) when = 0. Analogously to
the robust optimization problem discussed above, the dual of
the problem of maximizing

qQ
f
q

q
subject to the above
constraints gives a set of linear constraints that can replace the
nonlinear capacity constraint.
The continuous LP relaxation of the above problem is ob-
tained by replacing the last constraint with
q
[0, 1] q Q.
Let us use the set of dual variables
ij
,
q
ij
,
q
ij
,
q
ij
, and
ij
for
constraints (8)-(12) and
q
ij
for the upper bound constraint on

q
, i.e.,
q
1. As a consequence, in the robust optimization
model we must replace, for each ij E, the nonlinear
13
constraint (5) with the following set of constraints:
z
ij


qQ
v
avg
q
(
q
ij
+

q
ij
+
q
ij
) +K
ij
+

qQ

q
ij
ij E

ij

qQ
v
avg
q

q
ij
+

qQ
v
avg
q

q
ij
= 0 ij E

1
K

ij
+
ij
+
q
ij
v
avg
q
( 1)
q
ij

q
f
q
ij E, q Q

q
ij
+
q
ij
+
q
ij
= 0 ij E, q Q

q
ij
,
q
ij
,
q
ij
,
ij
,
q
ij
0 ij E, q Q.
Note that the larger set of constraints in problem P
ij
results
in a larger set of dual constraints and therefore a larger overall
robust optimization problem. This results in a MILP problem
that is similar in structure to that presented in Section III, but
more difcult to solve.
We have conducted some experiments using this model
as well, but only report a summary here for the sake of
conciseness. For = 0.1, i.e., an allowed reduction of 10%,
we have observed a decrease in the network cost between 13%
and 15% for the MPLS-only network, and a decrease of about
10% for the MPLS+OTN network. For = 0.05, instead, the
decrease is between 8% and 9% for the MPLS-only network
and 6% for the MPLS+OTN network.
F. Transponders and network cost
Even though we have so far considered node equipment
only, when factoring link costs in the model the comparison
becomes even more in favor of MPLS. In fact, one transponder
for each 10Gbps link has to be added to the network (see RFC
3031 [31]), both at the MPLS and at the OTN level.
Given that the cost of transponders is comparable to that
of 10 Gbps links and that there is no cost differential for
transponders used at either level, the effect will be that of a
smaller equivalent cost ratio between MPLS and OTN ports,
i.e., a smaller . As a consequence, the savings obtained from
an MPLS-only network or an optimized MPLS+OTN network
will be even larger than those discussed above.
V. CONCLUDING REMARKS
We have introduced a set of optimization models for the
allocation of capacity on networks with multiple layers. Tech-
nologies vary in cost and capability, and when more than one
are available it is difcult to choose one or a combination of
them that provides guaranteed service at a low cost.
Our models yield the most appropriate combination of
technologies and result in a network that serves multiple trafc
matrices. This class of models handles trafc uncertainty, and
computational experiments on a realistic network instance
show that they can provide a provably optimal solution in
reasonable computing times. Also, these models are able
to guarantee that all demands in a given uncertainty set,
characterized by simple parameters, are satised.
A. Open questions
There are several possible extensions to this work. First,
we have focused mostly on node equipment and considered
only node costs, but it might be of interest to investigate
further the effect of link costs on the network provisioning
process, for instance by explicitly taking transponder costs into
consideration.
Second, while OSPF weights are given as an input in our
problem, the network cost might benet from a scenario in
which the network operator that designs the network is also
allowed to set the OSPF weights. This would render the model
signicantly harder, but would also push optimization even
further. Previous attempts [32, 33] have shown that OSPF
weights have a signicant inuence on some network perfor-
mance indicators (congestion and routing cost respectively).
Third, it might be of interest to check whether relaxing
the requirement that all routing paths be shortest with respect
to OSPF weights further improves the network performance
(again, see [32]).
Partially related to this is the question whether, in networks
that have a high degree of ECMP (unlike the network in-
stance we have used), multiple shortest paths would make
a signicant difference in the network cost. In fact, in SP
networks, ECMP is very common, both for load balancing
and for resilience (if a link fails, having another path with the
same metric leads to less churn in the network). As a result,
every pair of edge routers has at least two paths, and often
many more.
Lastly, we have not considered the costs of adding fault
tolerance (or high-availability) in all models. Circuit networks
require end-to-end redundant conguration for managing fail-
ures (more than twice the cost of the non-redundant network).
MPLS networks allow for fast detection and reroute of trafc
during failure with discriminatory treatment for high priority
trafc, thus allowing for more efcient local link protection
and correspondingly lower costs.
REFERENCES
[1] D. Bienstock, S. Chopra, O. G unl uk, and C. Tsai,
Minimum cost capacity installation for multicommodity
network ows, Mathematical Programming, vol. 81, pp.
177199, 1998.
[2] L. S. Buriol, M. G. C. Resende, and M. Thorup, Surviv-
able IP network design with OSPF routing, Networks,
vol. 49, no. 1, pp. 5164, 2007.
[3] J. Kennington, E. Olinick, and G. Spiride, Basic math-
ematical programming models for capacity allocation in
mesh-based survivable networks, Omega, vol. 35, no. 6,
pp. 629644, 2007.
[4] T. Magnanti, P. Mirchandani, and R. Vachani, The
convex hull of two core capacitated network design prob-
lems, Mathematical Programming, vol. 60, pp. 233250,
1993.
[5] A. Ben-Tal, L. E. Ghaoui, and A. Nemirovski, Robust
Optimization. Princeton University Press, 2009.
[6] J. Birge and F. Louveaux, Introduction to stochastic
programming. New York: Springer Verlag, 1997.
[7] M. Dempster, E. Medova, and R. Thompson, A stochas-
tic programming approach to network planning, in the
15th International Teletrafc Congress (ITC), 1997, pp.
329339.
14
[8] A. Gaivoronski, Stochastic programming approach to
the network planning under uncertainty, in Optimization
in industry 3, A. Sciomachen, Ed. New York: John
Wiley and Sons, 1995, pp. 145163.
[9] S. Sen, R. Doverspike, and S. Cosares, Network plan-
ning with random demand, Telecommunication Systems,
vol. 3, no. 1, pp. 1130, 1994.
[10] J. Smith, A. Schaefer, and J. Yen, A stochastic integer
programming approach to solving a synchronous optical
network ring design problem, Networks, vol. 44, no. 1,
pp. 1226, 2004.
[11] S. Waller and A. Ziliaskopoulos, Stochastic dy-
namic network design problem, Transportation research
record, vol. 1771, pp. 106113, 2001.
[12] N. G. Dufeld, P. Goyal, A. Greenberg, P. Mishra, K. K.
Ramakrishnan, and J. van der Merive, A exible model
for resource management in virtual private networks,
in SIGCOMM 99: Proceedings of the conference on
Applications, technologies, architectures, and protocols
for computer communication, Cambridge, MA, 1999, pp.
95108.
[13] A. Gupta, J. Kleinberg, A. Kumar, R. Rastogi, and
B. Yener, Provisioning a virtual private network: A
network design problem for multicommodity ow, in
33rd ACM Symposium on Theory of Computing, 2001,
pp. 389398.
[14] F. Eisenbrand, F. Grandoni, F. Oriolo, and M. Skutella,
New approaches for virtual private network design,
SIAM J. Computing, vol. 37, no. 3, pp. 706721, 2007.
[15] A. Altn, E. Amaldi, P. Belotti, and M. C . Pnar, Provi-
sioning virtual private networks under trafc uncertainty,
Networks, vol. 49, no. 1, pp. 100115, 2007.
[16] W. Ben-Ameur and H. Kerivin, Routing of uncertain
trafc demands, Optimization and Engineering, vol. 6,
no. 3, pp. 283313, 2005.
[17] A. Ben-Tal and A. Nemirovski, Robust convex opti-
mization, Mathematics of Operations Research, vol. 23,
no. 4, pp. 769805, 1998.
[18] F. Alizadeh and D. Goldfarb, Second-order cone pro-
gramming, Mathematical Programming, vol. 95, no. 1,
pp. 351, 2003.
[19] D. Bertsimas and M. Sim, Robust discrete optimization
and network ows, Mathematical Programming, vol. 98,
no. 1, pp. 4971, 2003.
[20] P. Belotti, A. Capone, G. Carello, and F. Malucelli,
Multi-layer MPLS network design: The impact of sta-
tistical multiplexing, Computer Networks, vol. 52, no. 6,
pp. 12911307, 2008.
[21] M. Riis, A. Skriver, and S. Mller, Internet protocol
network design with uncertain demand, Journal of the
Operational Research Society, vol. 56, pp. 11841195,
2005.
[22] N. Springs, R. Mahajan, and D. Wetherall, Measuring
ISP topologies with Rocketfuel, IEEE/ACM Transac-
tions on Networking, vol. 12, no. 1, pp. 216, 2004.
[23] E. G. Coffman, Jr., M. R. Garey, and D. S. Johnson, Ap-
proximation algorithms for bin packing: a survey, in Ap-
proximation algorithms for NP-hard problems. Boston,
MA, USA: PWS Publishing Co., 1997, pp. 4693.
[24] M. R. Garey and D. S. Johnson, Computers and In-
tractability: A guide to the theory of NP-completeness.
W. H. Freeman and Co., 1979.
[25] G. L. Nemhauser and L. A. Wolsey, Integer and Combi-
natorial Optimization. John Wiley & Sons, 1988.
[26] L. N. Vicente and P. H. Calamai, Bilevel and multilevel
programming: A bibliography review, Journal of Global
Optimization, vol. 5, no. 3, pp. 291306, 1994.
[27] A. L. Soyster, Convex programming with set-inclusive
constraints and applications to inexact linear program-
ming, Operations Research, vol. 21, no. 5, pp. 1154
1157, 1973.
[28] V. Chv atal, Linear Programming. W. H. Freeman, 1983.
[29] G. Oriolo, Domination between trafc matrices, Mathe-
matics of Operations Research, vol. 33, no. 1, pp. 9196,
2008.
[30] R. Fourer, D. M. Gay, and B. Kernighan, AMPL: a
mathematical programming language, in Algorithms and
model formulations in mathematical programming. New
York, NY, USA: Springer-Verlag, 1989, pp. 150151.
[31] E. Rosen, A. Viswanathan, and R. Callon, RFC 3031:
Multiprotocol Label Switching Architecture, January
2001, available at http://www.ietf.org/rfc/rfc3031.txt.
[32] A. Altn, P. Belotti, and M. Pnar, OSPF routing
with optimal oblivious performance ratio under polyhe-
dral demand uncertainty, Optimization and Engineering,
vol. 11, no. 3, pp. 395422, 2010.
[33] B. Fortz and M. Thorup, Internet trafc engineering by
optimizing OSPF weights, in IEEE INFOCOM, vol. 2,
2000, pp. 519528.

You might also like