You are on page 1of 20

Discrete Rayleigh Fading Channel Modeling

Julio Aráuz
jarauz@mail.sis.pitt.edu
March 2002
Department of Information Sciences and Telecommunications
University of Pittsburgh
135 N. Bellefield Ave.
Pittsburgh PA, 15260

Abstract
Diverse methods that have been proposed to model Rayleigh fading channels will be
explored in this tutorial. Preliminary concepts related to wireless communications
channel modeling are identified before introducing the models. The first model presented
is the finite state Markov channel model which is based on the side information given by
the received signal to noise ratio. The Markovian validity of the model is described along
with the adequate conditions under which such validity holds. The relationships between
the model and modulation schemes, error control protocols and channel coding are also
described. Additionally hidden Markov models proposed to model Rayleigh fading
channels are also studied along with their validity assumptions.

1. Introduction
The field of mobile communications is changing the way people interact in their daily lives. The
wireless industry has developed and deployed an infrastructure that aims at providing diverse services for
the market. New wireless communications services offer people the possibility of being connected almost
anywhere they go. However the design, production and deployment of such technological infrastructure
come along with a high cost. This high cost may hinder manufacturers from building actual systems to
test their initial designs. Therefore manufacturers look at different alternatives to avoid high costs; one of
these alternatives is simulating a real wireless system. The advantage of simulations is that they could
allow less expensive testing of designs, although they could require previous investments on computing
resources.
A simulation of a wireless system could depend on various components. For example an end to
end simulation of a discrete system could include blocks to study voice encoding, channel coding,
interleaving and modulation issues. A key component of a wireless system simulation is the wireless
channel model. The modeling of the wireless channel in a simulation will dictate, for example, how bits
or packets are lost. These factors will definitely influence the overall performance of the system.

1
Different approaches could be used to simulate diverse types of channels and their conditions. For
example some models can be used to study and represent physical layer conditions like SNR or BER
while others may be used to study higher layer issues like packet or segment error rates as a function of
time.
Wireless channel models are commonly used to study, via simulations, the performance of
transport or link layer protocols. For example, Chaskar et al [9] study the performance of a transport
protocol (TCP) over wireless links; Chiani et al [10] perform an analytical study of TCP/IP over wireless
links. Labiod [15] studies the performance of error correcting codes over wireless links. In all of these
cases a model for the losses of frames at the link layer level has been implemented. Furthermore, the
performance of other communication protocols, such as ATM [12] [24], over wireless links has also been
studied via simulations. The validity of these performance studies is influenced by that of the underlying
channel model. The common use of channel models gives rise to questions about their validity. This
originates this tutorial’s motivation, to study these models and the conditions under which they can be
used.
This document presents a tutorial on current methods being used to model flat Rayleigh fading
channels. Since the models that will be described include a common basis of concepts, the tutorial starts
by identifying and describing them. After the basic concepts have been introduced a common model will
be presented, this is the Finite State Markov Channel (FSMC) model [36]. Based on this model, current
and past literature elaborate on the analysis of the model’s capacity, application areas and correctness;
therefore the main results of these analyses will be included. Along with this assessment, several points
will be highlighted by the author to illustrate where there could be a possibility for future work on the
topic.
Following the details on the general Markov modeling of Rayleigh fading wireless channels a
different approach will be introduced. This approach makes use of hidden Markov models (HMM) to
approximate the characteristic of the channel. Two key advantages of using HMM is that current
literature presents ‘powerful’ methods for fitting the models to experimental data [31] and that it is
possible to obtain closed solutions for several parameters of the model. Both of these advantages could
result in better approximation of the model to specific conditions.

2. Preliminary concepts
It is relevant to remember that the main goal of this tutorial is to present methods to model a
Rayleigh fading channel. Therefore it is important to briefly review the concepts behind fading. In a
wireless communication system the signals may travel through multiple paths between a transmitter and a

2
receiver [26]. This effect is called multipath propagation. Due to the multiple paths, the receiver of the
signal will observe variations of amplitude, phase and angle of arrival of the transmitted signal. These
variations originate the phenomenon referred as multipath fading. The variations are characterized by
two main manifestations [26], large-scale and small-scale fading. Furthermore these manifestations give
rise to specific types of degradations of the signal. Based on [26] figure 1 presents the fading
manifestations and its associated degradations.

Fading
Manifestations

Large Scale Small Scale


Fading Fading

Signal Time variance


dispersion of the channel

Degradations
Frequency Flat Fast Slow
selective fading Fading Fading
fading

Figure 1. Channel fading manifestations and degradations

The first fading manifestation, large-scale fading, refers to path loss caused by the effects of the
signal traveling over large areas. Large-scale fading characterizes the losses due to considerably big
physical objects in the signal’s path like hills or forests. The path loss is characterized [26] by a mean
loss (due to the distance between the transmitter and the receiver and the propagation environment
characteristics) and a variation around the mean loss.
On the other hand small-scale fading characterizes the effects of small changes in the separation
between a transmitter and a receiver. These changes can be caused by mobility of the transmitter,
receiver or the intermediate objects in the path of the signal. Small scale changes result in considerable
variations of signal amplitude and phase. Small-scale fading is also known as Rayleigh1 fading since the

−r2
r
1
The mathematical expression for the Rayleigh probability distribution is: p(r ) = e 2b for r ≥0 and p (r ) is zero for r<0.
b

3
fluctuation of the signal envelope is Rayleigh distributed when there is no predominant line of sight
between the transmitter and receiver. When there is a predominant line of sight between the transmitter
and receiver the fluctuations are statistically described by a Rician p.d.f [23].
Figure 1 shows two manifestations of small scale fading. The first one, signal dispersion, refers
to the time spreading of the signal. Dispersion causes the underlying digital pulses transmitted in the
signal to spread in time. The second manifestation reflects the time variant behavior of the channel that is
due to relative mobility between a transmitter and a receiver or the objects in the path of the signal. Both
of these manifestations can be characterized in the time and frequency domain by fading degradation
types.
As shown in figure 1, the degradation types of the dispersion manifestation are frequency
selective fading and flat fading. From the time domain point of view, frequency selective fading occurs
when the maximum spread in time of a symbol is greater than the duration of the symbol. Consequently,
another name for this fading degradation is channel induced intersymbol interference. From the
frequency domain point of view, frequency selective fading occurs when the spectral components of a
signal are affected in different ways by the channel. In particular, frequency selective fading occurs when
the channel’s coherence bandwidth (the channel’s bandwidth in which all components experience
approximately the same fading characteristics) is smaller than the signal’s bandwidth. When the
conditions described above, for frequency selective fading are not met, the degradation is referred as flat
fading. In this case the channel characteristics are approximately flat for all frequencies.
Figure 1 also shows the degradation types of the channel’s time variance manifestation. These
are fast and slow fading. From the time domain point of view, fast fading refers to the condition in which
the channel’s coherence time (an expected time duration during which the channel’s response is invariant)
is smaller than the symbol duration.
Before describing fast fading in the frequency domain it is necessary to introduce the Doppler
frequency concept. The Doppler frequency (fm) characterizes the maximum Doppler frequency shift of
the signals in a mobile environment. This is computed as f m = v / λ . Where ‘ v ’ is the relative velocity

between the transmitter and receiver and ‘λ’ is the wavelength of the transmitted signal. With this basis
one can indicate that from the frequency domain point of view, fast fading occurs when the signal
bandwidth is less than the maximum frequency Doppler shift.
Both types of small scale fading can be present in a wireless system. In this tutorial we will look
at flat fading with the associated time variance manifestation. Large scale fading is reflected only on the
strength of the received signal and will not be considered here. Having presented some relevant
preliminary information we can proceed to the central topic of this document, modeling of flat Rayleigh
fading channels.

4
3. Wireless channel models
Among the common models proposed to characterize a flat Rayleigh fading channel one can find
the Gilbert-Elliot channel, the FSMC [36] and hidden Markov models (HMM) [33]. The Gilbert-Elliot
channel is a special case of a FSMC; on the other hand HMMs use a different approach to model the
channel. The models in this section will be divided in FSMC models and HMMs.

3.1 Finite State Markov Channels


Wang and Moayeri [36] propose the modeling of a Rayleigh fading channel by the use of a
Markov process2 with a finite number of states referred to as the Finite State Markov Channel (FSMC)
model. The FSMC channel originated as an extension of a simpler model proposed earlier, the Gilbert-
Elliot channel model. This tutorial will first present the FSMC and then it will proceed to describe the
Gilbert-Elliot model. Following the approach of [36] this tutorial will start by establishing the general
basis for Markov modeling of wireless channels and then proceed to link this basis with specific physical
characteristics of the real channel. In order to avoid mathematical particularities, the actual mathematical
expressions of the model will not be developed but only described.
The theory behind the FSMC model is that of constant Markov processes. Constant Markov
processes have the property that the state transition probabilities are independent of the time at which they
occur. These processes can be defined by a finite number of possible states that are usually represented
by a set S= {so, s1, … , sn-1} and a sequence of states {Sk}, k=0,1,2….. Each of the elements of the set S
will later be associated with a characteristic of the channel; for now let’s just affirm that each state
represents a condition in which we have some probability of having an error in a symbol being
transmitted over the channel. For the moment, we are not linking this condition to a physical concept.
There are some basic particularities that should be established to fully describe the mathematical
model. Nevertheless, remember we are temporarily trying to detach the mathematical details from the
channel’s physical reality. The next table summarizes, as described in [36], what is necessary to
mathematically describe the FSMC, the reader should notice the terminology and notation used is the
same used for Markov processes. The table shows the elements of a ‘n’ state FSMC model.

2
A Markov process with a discrete state space is referred to as a Markov chain [5, pg. 21]

5
Component Notation Description for a ‘n’ state FSMC model
Transition P A n×n matrix representing the probability of transition between
probability matrix states or into the same current state.
Steady state π A 1×n vector representing the steady state probability of being in
probability vector any of the n states (remember that πP=P)
Crossover e A 1×n vector representing the different probabilities of have a
probability vector symbol in error in any of the n states

Table 1. Elements necessary to describe the Finite State Markov Channel (FSMC)

It is important to understand that with the elements defined Table 1, it is possible (under certain
assumptions that will be described later) to mathematically obtain relationships between each state, their
duration and its physical meaning [38].
The elements shown in Table 1 have some constraints. Any probability of the matrix P should be
between 0 and 1; the rows of P should add up to one and the elements of e should be between 0 and 0.5.
Having mathematically described the FSMC, let’s draw what the model proposes.
p01 p12

p11
p00
State0 State1 State2 State
n-1

p10 p21
p0 p1 p2 pn-1
0 0 0 0 0 0 0 0
BSCs
{ 1
1-p0

1 1
1-p1

1 1
1-p2

1 1
1-pn-1

Figure 2. The Finite State Markov Channel model representation

In figure 2 one can observe a FSMC with ‘n’ states. The figure shows the in-state transition
probabilities (pii), adjacent state transition probabilities (pij) and in its lower part the probability of a
symbol error in each state (1-pi). At this point we can make the first connection between this model and
the reality being modeled. This connection lies on the elements of the vector e and each state. The
elements of the vector e, called crossover probabilities, represent the crossover probabilities of ‘n’ binary
symmetric channels (BSC). The ‘n’ BSCs are drawn in the lower part of figure 1 and illustrate how a
symbol being transmitted, for example a zero or a one, can be received in error in any of the n states. As

6
shown in the figure a BSC is associated with each of the n states and its goal is to relate the varying nature
of the channel, and its error processes characteristics with the FSMC.
Of course there are additional restrictions over the concepts already introduced. Not any set of P,
π and e that satisfy the mathematical expressions for the FSMC [36] will adequately approximate the
model to a real physical channel. Therefore it is necessary to establish relationships between these
abstract elements and reality. Additionally, it is important to mention that the FSMC characterization will
be restricted to a discrete channel structure case, this is a digital modulation scheme will be assumed in
order to compute all the necessary values for the parameters.
The first relationship that must be drawn is that between the states and the actual channel. At this
point it is necessary to link each of the states to a physical concept, received signal to noise ratio (SNR).
The SNR is an important ‘side’ parameter that is used to represent a channel’s quality [37]. For example
given a specific modulation scheme it is possible to compute the probability of having a symbol in error
as a function of the received SNR [8]. Since the received signal envelope has a known probability
distribution (i.e. Rayleigh distributed) and the received SNR is proportional to the square of the signal
envelope one can compute the p.d.f of the received SNR. Reference [21] shows that the p.d.f of the SNR
in a Rayleigh fading environment with Gaussian noise is exponentially distributed. Based on this fact, the
FSMC model partitions the received SNR into a finite number of intervals (n) and represents each
interval as a state of the Markov process. The next figure shows a graph of a received signal to noise
ratio as a function of time at a receiver and how the partitioning could be performed.
Received SNR

Interval assigned to state n

Interval assigned to state 1

Interval assigned to state 0

time
Figure 3. Partitioning the received SNR and assigning each interval to a state of the FSMC

The partitioning in figure 3 has created ‘n’ SNR intervals, the first one starting at level of zero
SNR and the last one including all received SNR values greater than a certain threshold.
Wang and Moayeri [36] show that with the partition intervals defined (although no information
on how the partitioning should be done is given) and assuming BPSK modulation, it is possible to obtain
closed form expressions for the elements of the steady state probability vector π and the elements of the
crossover probability vector e. These relationships are important because they later on help in verifying

7
how closely they match the simulation results. In summary, up to this point, with the partitioning
information and the modulation scheme one can compute, as shown in detail in [36] and described later,
the values for the elements π and e.
For the model to be complete it is necessary to compute the values of the transition probabilities
between states (elements of the matrix P). As shown in [36], assuming that the fading is slow enough
helps in determining these probabilities, that way the received SNR remains constant during a channel
symbol. Under slow fading conditions the level crossing rate at any particular SNR partitioning level is
very small compared to the total time spent in that state. Reference [11] shows that given the maximum
Doppler frequency (fm) it is possible to compute the number of times (Na) that the received SNR passes
downward across a certain level. Additionally the symbols per second transmitted in any state (and thus
the total amount of time spent in a state) can be approximated with the elements of the vector π and the
transmission rate.
Following the procedure described above the symbols per second transmitted during the time
spent in state k, noted as R,k is equal to (transmission_rate × πk), Therefore the transition probabilities
can be approximated by the relation between the crossing rate and the rate symbol in each state. For
example the probability of going from state k to state k-1 is approximately equal to Nk/Rk. Reference
[36] verifies through simulations the appropriateness of these approximations.
As mentioned elsewhere the FSMC is a generalization of the Gilbert-Elliot channel model. This
model includes only two states. These two states are commonly referred as “good” and “bad” states. The
good state is called “noiseless” since the crossover probability of the associated binary symmetric channel
is zero. This means that during the good state no bits transmitted through the channel arrive with errors.
The “bad” state is called totally noisy since the crossover probability of the associated binary symmetric
channel is 0.5.
The following figure illustrates a Gilbert-Elliot channel model and its associated BSCs.
µ

State1 State2

λ
p1=0.5 p2=0
0 0 0 0
1-p1 1-p2

1 1 1 1

Figure 4 – Gilbert Elliot Channel as usually implemented in simulations

8
Figure 4 shows a common Gilbert-Elliot channel representation in which there are no in state
transitions. This model is usually used for simulations [9, 10, 12, 15, 18] because of the simplicity
associated with its implementation. It is important to mention that, for an actual implementation in
simulations the transition probabilities are not directly used, but the time spent in each state. In relation to
this fact, the arrows in the figure indicate the transition rates µ and λ between states. The mean time in
each state is computed based on the duration of a fade of the signal [11] in relation to an ‘acceptable
level’. Received signals that are below an acceptable level are modeled by the “bad” state. The “good”
state represents strong enough signals that are above a certain level.
Up to this point this document has presented the methods for modeling FSMC, but little has been
said about its validity, accuracy, appropriate partitioning intervals or relationship with modulation or
coding schemes. Current literature reviews these areas but also leaves some open questions about the
appropriate conditions under which the model is valid.

3.1.1 Validity and Accuracy of the FSMC


The author believes that the most important way of validating any model should consist in
comparing its results with experimental data. For example a comparison between the distributions of the
time spent in each state could be performed between the model and experimental data from a sample
function of the underlying random process. Previous literature mainly presents validations like this, but
by comparing simulated Rayleigh fading channels with the FSMC results [36]. Later results [5] also
show that experimental channel data appears to be suitable for modeling it with a Markov process. In
regards to validation, ther are two common approaches. One common type of validation that has been
thoroughly developed is based on an information theory analysis. The second one compares the
correlations of the processes under analysis. For both cases, as will be presented here, simulations have
been used in the literature to study the approaches.

Information Theory Validation Analysis


Before presenting the analysis for validating the FSMC, it is relevant to emphasize that the
proposed model conforms to the Markov property. In general, the Markov property can be expressed
as [13]:
p[ S (t n ) = s n | S (t n −1 ) = s n −1 , S (t n− 2 ) = s n −2 ,..., S (t o ) = s 0 ] = p[ S (t n ) = s n | S (t n −1 ) = s n −1 ]
This property indicates that the probability of transition at a time ‘n’ to a new state only depends on the
state at time ‘n-1’ (also referred to as first-order assumption). For the FSMC we are therefore assuming
that the ‘history’ of the previous channel states, besides the previous one, does not carry significant

9
information about the next state. Without any further analysis it is difficult to visualize if the Rayleigh
fading channel can be modeled following this assumption (later on we will describe why this assumption
is valid and under what conditions). Furthermore it could appear to be more desirable to have a model
that includes higher order assumptions and therefore maybe increase its accuracy [2, 3]. However, the
problem with higher order models is that the complexity of its analysis and implementation increases
considerably.
Wang and Chang [37] proposed a mutual information metric to verify the accuracy of the first
order Markovian assumption for a Rayleigh FSMC model. The goal of the metric is to confirm that given
the information about the previous symbol, the uncertainty of the current one should be negligible. This
uncertainty is measured in terms of average mutual information3 of the received amplitudes. Let’s
proceed to describe the common information theory analysis of FSMC.
Let Ai (where i is the time index) be the received SNR received amplitude of the ith symbol. The
information contained in Ai given by the two consecutive (and previous) amplitudes Ai-1Ai-2 is quantified
by the average mutual information I(Ai;Ai-1Ai-2). As proposed in [17], this can be expressed in terms of
the average conditional mutual information I(Ai;Ai-2|Ai-1) which can be expanded as follows.
I(Ai;Ai-1Ai-2) = I(Ai;Ai-1) + I(Ai;Ai-2|Ai-1).
Wang and Chang’s goal is to compute the value for the ratio I(Ai;Ai-2|Ai-1) / I(Ai;Ai-1Ai-2), which
as they show is a function of the joint p.d.f of Ai, Ai-1 and Ai-2. Additionally they show how this p.d.f
depends on the symbol transmission rate and other channel characteristics.
If the ratio I(Ai;Ai-2|Ai-1) / I(Ai;Ai-1Ai-2) is much smaller than one, then the average mutual
information I(Ai;Ai-1Ai-2) mainly depends on the first term, I(Ai;Ai-1). This would mean that the
information of Ai would mainly depend on the previous symbol Ai-1. If this happens, the first order
assumption for the FSMC would be verified. Since the joint p.d.f of Ai, Ai-1 and Ai-2 depends on physical
characteristics it is important to describe what these are and their ranges in order to maintain the FSMC
validity.
The results presented in [37] show that for the range of the product fm×τ (where 1/ τ is the
symbol transmission rate) going from 2×10-4 to 4×10-3 the value of the ratio I(Ai;Ai-2|Ai-1) / I(Ai;Ai-1Ai-2) is
less than 1%. This value is even smaller for small values of fm×τ, since as fading gets slower the
information of Ai is basically a function of Ai-1 only. Therefore using higher order models will not
improve the accuracy of the FSMC. On the other hand, for cases in which fast fading is observed the
value of the ratio indicates that this is not negligible and the first order assumption is not longer valid.
We will explore the actual limits for slow and fast fading validity of the FSMC later in this tutorial.

3
The average mutual information I(X;YZ) is a metric that indicates the amount of information about X that can be provided by YZ.

10
Stochastic Validation Analysis
With the results presented in [37] the accuracy of the first order model is verified but nevertheless
as pointed out by Tan and Beaulieu [30] the fact that one has “small mutual information is not a sufficient
condition to indicate a process is Markovian”. Reference [30] indicates that I(Ai;Ai-2|Ai-1) can approach
i, i-1 i-2
zero in two cases. The first case is when the samples at and are independent or when they are
highly correlated. Under very slow fading conditions, such as those explored in [37], this is the case.
Tan and Beaulieu [30] propose that an appropriate way of verifying the accuracy of the first order
Markovian assumption is to analyze I(Ai;Ai-1,Ai-2, Ai-3,…, A-∞). The original validation in which Wang
and Chang [37] analyze the value of I(Ai;Ai-2|Ai-1) only indicates that a second order Markovian model is
marginally better than a first order one, but does not indicate that even higher order models are not better
than the first order one. The tractability of the joint p.d.f motivates the usage of a different type of
analysis. Reference [30] proposes a second type of analysis, a “stochastic” analysis, of the FSMC.
In the stochastic analysis the autocorrelation functions of the FSMC4 model and a ISORA5
(isotropic scattering, omnidirectional receiving antenna) model are compared. The comparison of the
autocorrelation functions of these two models provides an insight on how well the FSMC matches a
generic ‘real’ model. The results presented in [30 Fig. 2,3] indicate that the autocorrelation of a FSMC in
general significantly differs from the ISORA model. These differences between the autocorrelation
functions are more noticeable as the fading speed increases. For example, at a normalized value of fmτ of
0.002 the two models appear to be more consistent with one another. While at a value of fmτ of 0.02 the
differences are quite noticeable. Additionally, from [30] it can be inferred that for slow fading conditions
the two autocorrelation functions tend to match each other as the number of states (n) of the FSMC
increases. Reference [30] uses a FSMC with a number of states that is in the range of 50 to 1000 states.
The autocorrelation functions shown by Tan and Beaulieu also suggests that first order Markov
chains, like the first order FSMC, are appropriate for very slowly fading channels but only for ‘very
slowly fading’ applications. Very slowly fading applications are those that require analysis over a short
duration of time. An example of a very slowly fading application could be the analysis of error
correction code block-error rates, which according to [30] requires analysis over a moderate number of
consecutive samples. By analyzing how the autocorrelation functions diverge over an increasing
separation between sampling points, [30] arrives to the conclusion of how the FSMC is valid for very

4
Reference 16 refers to the FSMC as AFSMC (Amplitude Based FSMC), in this tutorial FSMC will be used to maintain consistency.
5
The ISORA model refers to a complex Gaussian process with an envelope described by a stationary process with a Rayleigh first-order
distribution.

11
slowly fading applications. Tan and Beaulieu do not elaborate on the fact that these very slowly fading
applications if analyzed with the FSMC should also be analyzed under slow fading conditions in order for
the model to be valid. For very slowly fading applications, the functions of both the ISORA and the
FSMC are very similar for distinct values of sample separations [30 Fig. 5].
Having described the accuracy of a FSMC model under very slowly fading conditions we can
proceed to portray what happens under fast fading conditions. As mentioned before under fast fading the
autocorrelation function of the FSMC and the ISORA model differ but both tend to approach the
conditions of an uncorrelated model over any fixed sample separation. This leads to conclude that an
uncorrelated model is suitable under fast fading conditions. The details of how this model should be
formulated are not given in [30] although it is pointed out that it is implementation and analysis is much
simpler than that of the FSMC.

3.1.2 Partitioning, modulation and coding concerns in the FSMC model


Partitioning
Up to this point we have outlined the mathematical definitions for a FSMC model and related
them to real physical characteristics of a Rayleigh fading channel. We then proceeded to elaborate on the
validity and accuracy of a first order model and detailed under what conditions the model can be applied.
In order to construct the actual model one of the parameters that must be specified is how the partitioning
of the received SNR could be done.
Wang and Moayeri [36] completely describe the channel model but do not get involved with the
details of the partitioning of the SNR. In the literature a common approach for the partitioning is to select
the thresholds in such a way that the steady state probabilities of being in any state are equal. In terms of
the terminology used elsewhere this means that π0= π1 … πn-1=1/n (for a ‘n’ state FSMC). Right away one
could appreciate that the simplicity of this partitioning does not take into account the non linearity
between SNR and the individual symbol probabilities of the BSCs of the model. Therefore, some kind of
optimization can be performed.
Current literature does not elaborate extensively on how to select a partitioning scheme. An open
area to explore could be that of determining the effects on the model of placing a higher number of states
in the regions where the average SNR value falls more often. This of course will depend on the specific
average value of the SNR. Furthermore, the number of upward crossings through a certain level and fade
duration vary as a function of the SNR level [11, pg 35,37] therefore the effect of different partitioning
schemes that take this into consideration can also be investigated.

12
An optimization of how the partitioning can be done is proposed in [35]. In addition to the
simple partitioning method mentioned above, [35] suggests two other partitioning schemes for the SNR.
A first approach suggests to make πi = 2 πi-1 and π0 = 1/(2n-1). This way the probability of being in a
‘higher’ level doubles at each higher level. A third scheme that has been studied proposes to make πi = i
π0 and π0 = 2/(n2+n). The way how these specific partitioning schemes are obtained or how they are
helpful in approximating real channel conditions is not clearly specified in [35].
Reference [35] studies the impact of these different partitioning schemes by analyzing the
capacity of the FSMC. The capacity they compute is the average capacity, in bits per channel use, over
k = n −1
all ‘n’ states, this is ∑π
k =0
k (1 − h(ek )) . Where ‘h’ is the binary entropy function6. Under average error

probabilities ranging from 0.005 to 0.1 it was found that while keeping the number of states fixed, the
capacity differences between the two later schemes are minimal (less than 1%). Greater differences in
capacity are observed between the first and two other partitioning schemes mentioned in the previous
paragraph. The literature however does not mention any other comparisons between the real channel
behavior and the partitioning schemes. It is not specified in the available literature how the partitioning
schemes approximate the conditions of the real physical channel.
A different approach for computing the partitioning thresholds and the number of states is given
in [38]. In general one could select the SNR ranges to be large enough so that a received symbol (or
packet) is completely received during the associated state. On the other hand, it is also desirable that
within a packet’s duration, similar SNR will be experienced so that the packet will experience similar
BER conditions, which are similar to the BER of the state. It is straightforward to understand that under
these guidelines the average duration of each state is a critical design parameter in order to make a one
step transition in the model after one packet time period. Zhang and Kassam [38] formulate a system of
linear equations that compute the states’ duration as a multiple of the duration of a packet (in particular
each state is assigned equal average duration). By solving a system of equations developed in [38] one
obtains the values for the SNR partitioning thresholds for a given number of states. Reference [38] does
not elaborate on what is an appropriate number of states. There are other references that have studied the
behavior of a FSMC with different number of states.
Babich and Lombardi [5] studied a two threshold (i.e. three states) FSMC model in a quantized
Rayleigh fading environment. They show that a first-order FSMC with three states gives a good
approximation of the fading process under sufficiently slow fading (fm×τ < 0.02). Under fast fading

6
The binary entropy h(.) as a function of the average error probability is defined as:
n −1
1
h(e) = e log + (1 − e) log
1 , where e = ∑ pk ek .
e 1− e k =0

13
conditions (fm×τ > 0.4), an uncorrelated model (zero-order model) proved to adequately approximate the
fading process. For intermediate values of fading the authors in [6] suggest a higher order model.
Nevertheless as mentioned before, current literature does not elaborate extensively on the selection
process of the number of states.
Up to this point we have detailed those results based on first order FSMC models. As pointed out
by [6], for intermediate fading conditions a higher order model could be more suitable to describe the
fading process. In reference [7], Babich et al propose a method for fitting Markov models of unspecified
order to narrow-band fading channels with additive Gaussian noise. The proposed method uses a context
tree pruning7 [3] algorithm to fit experimental or simulated fading sequences to a Markov model.

Modulation and coding


The parameterization of the FSMC model requires the specification of the symbol error
probabilities for the associated BSCs to each state. Since given a digital modulation scheme, the average
error probability is a function of the received SNR it is possible to compute crossover probabilities of
each associated BSC. In general fairly simply modulation schemes have been used in the study of FSMC
models. For example [36] studies the behavior of the model using BPSK, while other use π/4 DQPSK
[38]. No further relationships between the model and the modulation schemes are taken into
consideration in the parameterization.
How modulation influences a function of the fading process (statistics of block errors) is also
relevant, especially in simulations. Zorzi and Rao [39, 40] study the statistics of block errors when
transmitting data over fading channels. In [39] it was found that a simple two state first order Markov
model that describes the success/failure of transmitted blocks gives results that agree with those from a
detailed simulation (under slow fading conditions). Reference [40] studies the appropriateness of a
Markov approximation for the block error process. This modeling of the error process is necessary when
one studies the performance of upper layer protocols. In [40] the block error process is studied with
binary Markov models both at the block (threshold model) and symbol level. At the block level, a block is
considered in error when the value of the fading envelope is below a certain threshold as done in [39]. At
the symbol level, a symbol is considered in error with certain a probability that depends both on the
modulation scheme and the average SNR.
Additionally assuming in [40] the effect of using a (N,k) block code is studied. In the model, a
block is assumed to be correctly received when it contains fewer errors than those that the code is able
correct. By tracking the fading envelope at the symbol level the authors tried to include the effects of a

7
The Context Tree Pruning CTP algorithm estimates the best-fitting Markov model to a discrete sequence. According to [36] is optimal in the
sense that it provides the smallest parameterization.

14
varying envelope during the transmission of a block. The authors do not elaborate on the effect of
interleaving. Interleaving will have an effect on the error process since it spreads the errors and could
allow a more effective protection against burst of errors.
Using two modulation schemes, BPSK and FSK the authors in [40] investigate the sensitivity of
the block error process to coding and modulation schemes. They show that a Markov approximation for
the block error process is a ‘very good model for a broad range of parameters’. For example, for block
sizes (ranging from 100 to 2000 symbols per block), several error correcting capabilities and distinct
levels of modulations the authors show [40 Fig. 5, 6] that the threshold model greatly approximates the
results obtained by the symbol by symbol tracking process. These results indicate that for tracking
error processes at the block level under very slow fading conditions it is sufficient to use the two
state first order Markov model, this is the threshold model. The threshold model proves to be only
sensitive to the normalized Doppler frequency.
The following table summarizes the main approaches and proposals of different authors in
regards to the FSMC models.
Authors Type of study Relevant results, comments States, Partitioning used
[Reference]
Wang and FSMC mathematical Verified the accuracy of the 8 state model,
Moayeri [36] characterization and mathematical expressions for matrix P Equal probabilities
simulation with simulated results π0=π1…πn-1=1/n
st
Wang and Validation using Show how a 1 order FSMC is Not applicable
Chang [37] nd
information theory marginally better than a 2 order one.
analysis
Tan and Validation using Indicate that results from [37] are not 50 through 100 states,
Bealieu [30] information theory enough to prove the process is Equal probabilities
analysis and Markovian. Determine the ranges and
stochastic analysis applications of a first order FSMC
Wang and Capacity modeling of Show the impact of the number of 2 through 16 states,
Moayeri [35] the FSMC with states on the capacity and the effect of 1. Equal probabilities
different number of different partitioning schemes on the 2. πi = 2 πi-1 and π0 = 1/(2n-1).
states and partitions capacity
3. πi = i π0 and π0 = 2/(n2+n).
Zang and Methodology for Does not elaborate on the fact that SNR 2 through 16 states
Kassam [38] partitioning the SNR and e have a non linear relationship Equal probability method
Quantization method
Zorzi and Block error process: Threshold model approximates the 2 states
Rao [40] Symbol by symbol symbol by symbol model over a wide
tracking vs. Threshold range of parameters
model
Table 2. Summary of FSMC proposals

3.2 Hidden Markov Models


As previously mentioned the application of first order FSMC is adequate under very slowly
fading applications. Whenever there is a need to include the effect of very long channel memory the
FSMC model is not longer appropriate. This is the case of a study of fade duration distributions in fading

15
channels [34]. In this case there is a need of Markov chains with larger memory, but since the number of
states grows exponentially with the process memory [34] the approach is no longer practical. In such
cases other methods such as those that use Hidden Markov Models can be used.
Hidden Markov Models (HMM) [22] are probabilistic functions of Markov chains (also known as
Markov sources). These models can be used to study the fading process of a Rayleigh fading channel.
We will first start by defining the general characteristics and concepts related to HMMs. Then the tutorial
will proceed to describe how they are used to model fading.
A common discrete Markov process, like the one used in FSMC, is a stochastic process in which
the outputs are observable. The outputs in this case are the set of states at each instant of time.
Additionally each state corresponds to some physical and observable event. An ‘observable’ Markov
model could be too restrictive for some applications [22]. The observable models can be extended to
include the case where the ‘observation is a probabilistic function of the state’. This results in a ‘doubly
embedded stochastic process’ called Hidden Markov Model. One of the stochastic process is not
observable, hence the name Hidden Markov Model.

3.2.1 Characterization of HMMs


A HMM is characterized with the following elements.
1. The number of states ‘n’ in the model. It is important to note that even though the states are
‘hidden’ in practical applications these are associated with some physical event. The set of
the Markov chain states can therefore be represented by a set S = {s1, s2, … , sn}
2. The discrete alphabet size ‘m’. The alphabet corresponds to the set of outputs of the model in
any given state. The set X of the outputs can therefore be represented by as set X = {x1, x2,
…, xm}
3. The state transition probability distribution matrix A = {aij}, where aij = p[sj|si].
4. The observed symbol probability distribution matrix B. B is a diagonal matrix whose
elements bj represent the probability p{x | sj}, x є X. (if X is discrete).
5. The initial state probability vector π.
This tutorial does not have as a goal to fully describe the characterizations of HMMs, but to relate
these models to fading processes. As it will be explained later this relationship is similar to that of
FSMC models; here it will also be necessary to partition the received SNR and assign states to the
partitions. With the model already established, reference [34] shows how to compute the autocorrelation
functions and other statistics of HMM. We will elaborate more on this later.
In this tutorial we are interested in describing the methods that can be used to fit a HMM to a
specific fading process. The first fitting method than can be used is the method of moments [34]. In this

16
method the parameters of the model are found by equating the moments of the two models (i.e. HMM and
ISORA for example). This method has the problem that its system of equations is ‘ill posed’. This means
that the moments are the same for very different models. Additionally the selection of moments,
according to [34], is in general arbitrary. For example finding a HMM with an autocorrelation function
that resembles that of the fading process does not guarantees that the multidimensional probabilities
associated with these processes are close. The method of the moments is generally used to obtain a first
approximation that will be refined later with more advanced statistical methods.
A second fitting method consists in approximating multidimensional probability densities [34].
This method tries to answer the question of how to adjust the model parameters in order to maximize the
probability of having a certain observable sequence. In more specific terms, if the observation sequence
O is given, O= O1, O2, … , OT how do we best describe it? (Based on the model’s parameters). This
means we are trying to maximize the probability p(O|λ), where the model is λ = (A, B, π). The
observation sequence used to compute the model parameters is called training sequence.
There is no absolute optimal manner of estimating the model parameters to solve for the second
model fitting method. However there are methods to locally maximize p(O|λ) using an iterative
procedure. One of these iterative procedures is the Baum-Welch method (also called EM, expectation
maximization method) [22]. Reference [34] details several procedures that can be applied to optimize
the computational efficiency of the method. Additionally, an advantage of HMM modeling of fading
process is that it provides means to compute closed-form expressions for fade duration and level crossing
number distributions [34]. These expressions could be useful in the implementation of simulations.
Up to this point the actual HMM parameters have not been related to any real physical
characteristic of the fading channel. References [28] and [34] illustrate how this is done. The physical
reality is again connected to the HMM via the set of states S. As in the FSMC the fading amplitude needs
to be quantized and an element of the set S is assigned to each quantization level. In the references there
are no guidelines on how to select the threshold levels.
Via simulation of the fading envelope it is possible to compute the transition probability matrix
A. In a similar way the probability of the outputs of the model, this is matrix B, can computed. After
this, [34] proceeds to compute the state duration distribution of a Rayleigh fading channel using the
Baum-Welch algorithm. As shown in [34 Fig. 1] the approximation of the state distribution closely
resembles that obtained from simulation. In a similar manner, and as mentioned before, the fade duration
distribution and level crossing number distributions are also computed in [34] and their validity is
compared against simulated data.

17
The advantage of using HMMs is that they provide enough flexibility to model different types of
fading [32] [34]. Additionally if fading is modeled with a HMM then bit errors and block errors
occurring over fading channels can also be modeled with HMMs [31] [32].

4. Discussion
Throughout this tutorial we have analyzed two main approaches for the modeling of discrete
Rayleigh fading channels. These approaches are the FSMC channel and the HMMs. Several open issues
have been highlighted, among these, how current literature does not detail how to optimize the model in
regards to the number of states. Another interesting issue is that of partitioning the received SNR. In this
tutorial several partitioning approaches were introduced. The criteria to determine which partitioning
approach is better and under what circumstances have not been completely developed.
The two open topics mentioned before are relevant when a practical implementation of the model
is necessary. The reader should remember that for performance studies of higher level protocols the
simulation of the Rayleigh fading process is not the main issue, but the generation of a frame error
process associated with the underlying fading one. In current literature real physical channels are not
thoroughly compared with those obtained from simulation. Therefore the tuning criteria are not well
understood. One exception is the study presented in [5] where the order of the model and the number of
quantization levels are already given, but no detailed relationships on how the number of partitions affects
the performance of the model are given.
The concepts introduced in this tutorial have carried along an assumption that has not been
extensively mentioned. This is that of flat fading. As described in the introduction, throughout the
tutorial we have assumed that the fading is frequency non selective. This is not necessarily the case,
especially if the bandwidth used by the signal is big enough like in the case of spread spectrum signals.
For example it is not well understood how the models presented in the tutorial would describe the frame
error process of the wireless LAN signals like those from the IEEE 802.11b standard. Reference [14]
deals with the characterization of frequency selective Rayleigh fading channels, nevertheless comparisons
with real physical channels and tuning guidelines are not thoroughly developed.

REFERENCES
[1] A. Abdi, Correspondence from the IEEE Transactions on Vehicular Technology, Vol. 48, No. 5,
September 1999, p. 1739.
[2] F. Babich, G. Lombardi, “A Markov Model for the Mobile Propagation Channel”, IEEE
Transactions on Vehicular Technology, Vol. 49, No. 1, January 2000, pp. 63-73.
[3] F. Babich, G. Lombardi, “On Verifying a First-Order Markovian Model for the Multi-Threshold
Success/Failure Process for Rayleigh Channel”, PIMRC 97, Vol. 1, 1997, pp. 12-16.

18
[4] F. Babich, G. Lombardi, “Statistical Analysis and Characterization of the Indoor Propagation
Channel”, IEEE Transactions on Communications, Vol. 48, No. 3, March 2000, pp. 455-464.
[5] F. Babich, G. Lombardi, “A Measurement Based Markov Model for the Indoor Propagation
Channel”, IEEE 47th Vehicular Technology Conference, Vol. 1, 1997, pp. 77-81.
[6] F. Babich, O. Kelly, G. Lombardi, “Generalized Markov Modeling for Flat Fading”, IEEE
Transactions on Communications”, Vol. 48, No. 4, April 2000, pp. 547-551.
[7] F. Babich, O. Kelly, G. Lombardi, “A Context-Tree Based Model for Quantized Fading”, IEEE
Communications Letters, Vol. 3, No. 2, February 1999, pp. 46-48.
[8] A. Bateman, “A general analysis of bit error probability for reference based BPSK mobile data
transmissions”, IEEE Transactions on Communications, Vol. 37, April 1989.
[9] H. M. Chaskar, T.V. Lakshman, U. Madhow, "TCP Over Wireless with Link Level Error Control
Analysis and Design Methodology", IEEE/ACM Transactions on Networking, Vol. 7, No. 5,
October 1999. pp. 605-615.
[10] M. Chiani, E. Milani, Verdone R., “A semi-analytical approach for performance evaluation of TCP-
IP based mobile radio links”, IEEE Global Telecommunications Conference, 2000, Vol. 2, 2000,
pp. 937-942.
[11] W.C. Jakes, “Microwave Mobile Communications”, New York, McGraw Hill, 1989.
[12] J. G. Kim, M. Krunz, “Quality of service over wireless ATM links”, INFOCOM '99, Proceeding of
the IEEE Eighteenth Annual Joint Conference of the IEEE Computer and Communications
Societies, Vol. 3, 1999, pp. 1003-1010.
[13] L. Kleinrock, “Queueing Systems”, Vol. 1, New York, John Wiley & Sons, 1975.
[14] H. Kong, E. Shwedyk, “Markov Characterization of Frequency Selective Rayleigh Fading
Channels”, Proceedings of the IEEE Pacific Rim Conference on Communications, Computers, and
Signal Processing, 1995, pp. 359-362.
[15] H. Labiod, “Performance of Reed Solomon error-correcting codes on fading channels” IEEE
International Conference on Personal Wireless Communication, 1999, pp. 259-263.
[16] W. C. Y. Lee, “Mobile Communications Engineering”, New York, McGraw Hill, 1992.
[17] R. G. Gallager, “Information Theory and Reliable Communication”, New York, Wiley, 1968.
[18] J. Gómez, A. Campbell, “A Channel Predictor for Wireless Packet Network”, IEEE International
Conference on Multimedia and Expo, ICME 2000, Vol. 3, 2000, pp. 1269-1272.
[19] M. Mushkin, I. Bar-David, “Capacity and Coding for the Gilbert-Elliot Channels”, IEEE
Transactions on Information Theory, Vol. 35, No. 6, November 1989, pp. 1277-1289.
[20] N. Nefedov, “Generative Markov Models for Discrete Channel Modeling”, The 8th IEEE
International Symposium on Personal, Indoor and Mobile Radio Communications, 1997, Waves of
the Year 2000. PIMRC '97, Vol. 1, 1997, pp. 7-11.
[21] J. G. Proakis, “Digital Communications”, 2nd. Ed, New York, McGraw-Hill, 1989.
[22] L. R. Rabiner, “A Tutorial on Hidden Markov Models and Selected Applications in Speech
Recognition”, Proceedings of the IEEE, Vol. 77, No. 2, February 1989, pp. 257-286.
[23] T. Rappaport, “Wireless Communications”, Upper Saddle River N.J, Prentice Hall, 1996.
[24] C. Schuler, “Error correction strategies for wireless ATM”, The Fourth IEEE Workshop on High-
Performance Communication Systems, 1997. (HPCS '97), pp. 204-213.
[25] S. Sivaprakasam, K. S. Shanmugan, “An Equivalent Markov Model for Burst Errors in Digital
Channels”, IEEE Transactions on Communications, Vol. 43, No. 2/3/4, February/March/April
1995, pp. 1347-1355.
[26] B. Sklar, “Rayleigh fading channels in mobile digital communication systems part I:
Characterization”, IEEE Communications Magazine, Vol. 35, Issue 7, July 1997, pp. 90-100.
[27] B. Sklar, “Rayleigh fading channels in mobile digital communication systems part II: Mitigation”,
IEEE Communications Magazine, Vol. 35, Issue 7, July 1997, pp. 102-109.
[28] F. Swarts, H. C. Ferreira, “Markov Characterization of Channels with Soft Decision Outputs”,
IEEE Transactions on Communications, Vol. 41, No. 5, May 1993, pp.678-682.

19
[29] J. Swarts, H. Ferreira, “On The Evaluation and Application of Markov Channel Models in Wireless
Communications”, IEEE Vehicular Technology Conference, Vol. 1, Fall 1999, pp. 117-121.
[30] C. C. Tan, N. C. Beaulieu, “On First-Order Markov Modeling for the Rayleigh Fading Channel”,
IEEE Transactions on Communications, Vol. 48, No. 12, December 2000, pp. 2032-2040.
[31] W. Turin, “Performance Analysis of Wireless Systems using Hidden Markov Models”, IEEE
Vehicular Technology Conference Fall 2001 Tutorial.
[32] W. Turin, “Digital Transmission Systems”, New York, McGraw-Hill, 1999.
[33] W. Turin, M. M. Sondhi, “Modeling Error Sources in Digital Channels”, IEEE Journal on Selected
Areas in Communications, Vol. 11, No. 3, April 1993, pp. 340-347.
[34] W. Turin, R. Van Nobelen, “Hidden Markov Modeling of Flat-Fading Channels”, IEEE Journal on
Selected Areas in Communications, Vol. 16, No. 9, December 1998, pp. 1234-1238.
[35] H. S. Wang, N. Moayeri, “Modeling, Capacity and Joint Source/Channel Coding or Rayleigh
Fading Channels”, 43rd IEEE Vehicular Technology Conference, 1993, pp. 473-479.
[36] H. S. Wang, N. Moayeri, “Finite-State Markov Channel – A useful Model for Radio
Communications Channels”, IEEE Transactions on Vehicular Technology, Vol. 44, No. 1,
Februrary 1995, pp. 163-171.
[37] H. S. Wang, P. Chang, “On Verifying the First-Order Markovian Assumption for a Rayleigh Fading
Channel Model”, IEEE Transactions on Vehicular Technology, Vol. 45, No. 2, May 1996, pp. 353-
357.
[38] Q. Zhang, S. A. Kassam, “Finite-State Markov Model for Rayleigh Fading Channels”, IEEE
Transactions on Communications, Vol. 47, No. 11, November 1999, pp. 1688-1692.
[39] M. Zorzi, R. R. Rao, L. B. Milstein, “On the accuracy of a first-order Markov model for data
transmission on fading channels”, Fourth IEEE International Conference on Universal Personal
Communications, 1995, pp. 211-215.
[40] M. Zorzi, R. R. Rao, L. B. Milstein, “Error Statistics in Data Transmission over Fading Channels”,
IEEE Transactions on Communications, Vol. 46, No. 11, November 1998, pp. 1468-1476.
[41] M. Zorzi, R. R. Rao, “ARQ Error Control for Delay-Constrained Communications on Short-Range
Burst-Error Channels”, IEEE 47th Vehicular Technology Conference, Vol. 3, 1997, pp. 1528-1532.
[42] M. Zorzi, R. R. Rao, “Throughput analysis of Go-Back-N ARQ in Markov channels with unreliable
feedback”, IEEE International Conference on Communications, 1995. ICC '95 Seattle, 'Gateway to
Globalization', Vol. 2, 1995, pp. 1232-1237.
[43] M. Zorzi, R. R. Rao, “On the Statistics of Block Errors in Bursty Channels”, IEEE Transactions on
Communications, Vol. 45, No. 6, June 1997, pp. 660-666.
[44] Wireless and Mobility Extensions to ns The CMU Monarch Project,
http://www.monarch.cs.cmu.edu/, August 1999.

20

You might also like