History of network traffic models

Last updated

Design of robust and reliable networks and network services relies on an understanding of the traffic characteristics of the network. Throughout history, different models of network traffic have been developed and used for evaluating existing and proposed networks and services.

Contents

Demands on computer networks are not entirely predictable. Performance modeling is necessary for deciding the quality of service (QoS) level. Performance models in turn, require accurate traffic models that have the ability to capture the statistical characteristics of the actual traffic on the network. Many traffic models have been developed based on traffic measurement data. If the underlying traffic models do not efficiently capture the characteristics of the actual traffic, the result may be the under-estimation or over-estimation of the performance of the network. This impairs the design of the network. Traffic models are hence, a core component of any performance evaluation of networks and they need to be very accurate.

“Teletraffic theory is the application of mathematics to the measurement, modeling, and control of traffic in telecommunications networks. [1] The aim of traffic modeling is to find stochastic processes to represent the behavior of traffic. Working at the Copenhagen Telephone Company in the 1910s, A. K. Erlang famously characterized telephone traffic at the call level by certain probability distributions for arrivals of new calls and their holding times. Erlang applied the traffic models to estimate the telephone switch capacity needed to achieve a given call blocking probability. The Erlang blocking formulas had tremendous practical interest for public carriers because telephone facilities (switching and transmission) involved considerable investments. Over several decades, Erlang’s work stimulated the use of queuing theory, and applied probability in general, to engineer the public switched telephone network. Teletraffic theory for packet networks has seen considerable progress in recent decades. [2] [3] [4] [5] Significant advances have been made in long-range dependence, wavelet, and multifractal approaches. At the same time, traffic modeling continues to be challenged by evolving network technologies and new multimedia applications. For example, wireless technologies allow greater mobility of users. Mobility must be an additional consideration for modeling traffic in wireless networks. [6] [7] Traffic modeling is an ongoing process without a real end. Traffic models represent our best current understanding of traffic behavior, but our understanding will change and grow over time.” [8]

Network traffic models usage

Measurements are useful and necessary for verifying the actual network performance. However, measurements do not have the level of abstraction that makes traffic models useful. Traffic models can be used for hypothetical problem solving whereas traffic measurements only reflect current reality. In probabilistic terms, a traffic trace is a realization of a random process, whereas a traffic model is a random process. Thus, traffic models have universality. A traffic trace gives insight about a particular traffic source, but a traffic model gives insight about all traffic sources of that type. Traffic models have three major uses. One important use of traffic models is to properly dimension network resources for a target level of QoS. It was mentioned earlier that Erlang developed models of voice calls to estimate telephone switch capacity to achieve a target call blocking probability. Similarly, models of packet traffic are needed to estimate the bandwidth and buffer resources to provide acceptable packet delays and packet loss probability. Knowledge of the average traffic rate is not sufficient. It is known from queuing theory that queue lengths increase with the variability of traffic. [9] Hence, an understanding of traffic burstiness or variability is needed to determine sufficient buffer sizes at nodes and link capacities. [10] A second important use of traffic models is to verify network performance under specific traffic controls. For example, given a packet scheduling algorithm, it would be possible to evaluate the network performance resulting from different traffic scenarios. For another example, a popular area of research is new improvements to the TCP congestion avoidance algorithm. It is critical that any algorithm is stable and allows multiple hosts to share bandwidth fairly, while sustaining a high throughput. Effective evaluation of the stability, fairness, and throughput of new algorithms would not be possible without realistic source models. A third important use of traffic models is admission control. In particular, connection oriented networks such as ATM depends on admission control to block new connections to maintain QOS guarantees. A simple admission strategy could be based on the peak rate of a new connection; a new connection is admitted if the available bandwidth is greater than the peak rate. However, that strategy would be overly conservative because a variable bit-rate connection may need significantly less bandwidth than its peak rate. A more sophisticated admission strategy is based on effective bandwidths. [11] The source traffic behavior is translated into an effective bandwidth between the peak rate and average rate, which is the specific amount of bandwidth required to meet a given QoS constraint. The effective bandwidth depends on the variability of the source. [8]

Network traffic models steps

Traffic modeling consists of three steps:

Parameter estimation is based on a set of statistics (e.g. mean, variance, density function or auto covariance function, multifractal characteristics) that are measured or calculated from observed data. The set of statistics used in the inference process depends on the impact they may have in the main performance metrics of interest. [12]

Network traffic models parameter

In recent years several types of traffic behavior, that can have significant impact on network performance, were discovered: long-range dependence, self-similarity and, more recently, multifractality. There are two major parameters generated by network traffic models: packet length distributions and packet inter-arrival distributions. Other parameters, such as routes, distribution of destinations, etc., are of less importance. Simulations that use traces generated by network traffic models usually examine a single node in the network, such as a router or switch; factors that depend on specific network topologies or routing information are specific to those topologies and simulations. [13] The problem of packet size distribution is fairly well-understood today. Existing models of packet sizes have proven to be valid and simple. Most packet size models do not consider the problem of order in packet sizes. For example, a TCP datagram in one direction is likely to be followed by a tiny ACK in the other direction about half of one Round-Trip Time (RTT) later. The problem of packet inter-arrival distribution is much more difficult. Understanding of network traffic has evolved significantly over the years, leading to a series of evolutions in network traffic models.

Self-similar traffic models

One of the earliest objections to self-similar traffic models was the difficulty in mathematical analysis. Existing self-similar models could not be used in conventional queuing models. This limitation was rapidly overturned and workable models were constructed. Once basic self-similar models became feasible, the traffic modeling community settled into the “detail” concerns. TCP’s congestion control algorithm complicated the matter of modeling traffic, so solutions needed to be created. Parameter estimation of self-similar models was always difficult, and recent research addresses ways to model network traffic without fully understanding it. [14]

Ilkka Norros Let us start from scratch.jpg
Ilkka Norros

When self-similar traffic models were first introduced, there were no efficient, analytically tractable processes to generate the models. Ilkka Norros devised a stochastic process for a storage model with self-similar input and constant bit-rate output. While this initial model was continuous rather than discrete, the model was effective, simple, and attractive. [14]

All self-similar traffic models suffer from one significant drawback: estimating the self-similarity parameters from real network traffic requires huge amounts of data and takes extended computation. The most modern method, wavelet multi-resolution analysis, is more efficient, but still very costly. This is undesirable in a traffic model. SWING uses a surprisingly simple model for the network traffic analysis and generation. The model examines characteristics of users, Request-Response Exchanges (RREs), connections, individual packets, and the overall network. No attempt is made to analyze self-similarity characteristics; any self-similarity in the generated traffic comes naturally from the aggregation of many ON/OFF sources. [14] [15]

The Pareto distribution process produces independent and identically distributed (IID) inter-arrival times. In general if X is a random variable with a Pareto distribution, then the probability that X is greater than some number x is given by P(X > x) = (x/x_m)-k for all x ≥ x_m where k is a positive parameter and x_m is the minimum possible value of Xi The probability distribution and the density functions are represented as: F(t) = 1 – (α/t)β where α,β ≥ 0 & t ≥ α f(t) = βαβ t-β-1 The parameters β and α are the shape and location parameters, respectively. The Pareto distribution is applied to model self-similar arrival in packet traffic. It is also referred to as double exponential, power law distribution. Other important characteristics of the model are that the Pareto distribution has infinite variance, when β ≥ 2 and achieves infinite mean, when β ≤ 1.

The Weibull distributed process is heavy-tailed and can model the fixed rate in ON period and ON/OFF period lengths, when producing self-similar traffic by multiplexing ON/OFF sources. The distribution function in this case is given by: F(t) = 1 – e-(t/β)α t > 0 and the density function of the weibull distribution is given as: f(t) = αβ-α tα-1 e -(t/β)α t > 0 where parameters β ≥ 0 and α > 0 are the scale and location parameters respectively. The Weibull distribution is close to a normal distribution. For β ≤ 1 the density function of the distribution is L shaped and for values of β > 1, it is bell shaped. This distribution gives a failure rate increasing with time. For β > 1, the failure rate decreases with time. At, β = 1, the failure rate is constant and the lifetimes are exponentially distributed.

The Autoregressive model is one of a group of linear prediction formulas that attempt to predict an output y_n of a system based on previous set of outputs {y_k} where k < n and inputs x_n and {x_k} where k < n. There exist minor changes in the way the predictions are computed based on which, several variations of the model are developed. Basically, when the model depends only on the previous outputs of the system, it is referred to as an auto-regressive model. It is referred to as a Moving Average Model (MAM), if it depends on only the inputs to the system. Finally, Autoregressive-Moving Average models are those that depend both on the inputs and the outputs, for prediction of current output. Autoregressive model of order p, denoted as AR(p), has the following form: Xt = R1 Xt-1 + R2 Xt-2 + ... + Rp Xt-p + Wt where Wt is the white noise, Ri are real numbers and Xt are prescribed correlated random numbers. The auto-correlation function of the AR(p) process consists of damped sine waves depending on whether the roots (solutions) of the model are real or imaginary. Discrete Autoregressive Model of order p, denoted as DAR(p), generates a stationary sequence of discrete random variables with a probability distribution and with an auto-correlation structure similar to that of the Autoregressive model of order p.[3]

Regression models define explicitly the next random variable in the sequence by previous ones within a specified time window and a moving average of a white noise.[5]

Transform-expand-sample (TES) models are non-linear regression models with modulo-1 arithmetic. They aim to capture both auto-correlation and marginal distribution of empirical data. TES models consist of two major TES processes: TES+ and TES–. TES+ produces a sequence which has positive correlation at lag 1, while TES– produces a negative correlation at lag 1. [16]

Non-self-similar traffic models

Early traffic models were derived from telecommunications models and focused on simplicity of analysis. They generally operated under the assumption that aggregating traffic from a large number of sources tended to smooth out bursts; that burstiness decreased as the number of traffic sources increased. [14]

One of the most widely used and oldest traffic models is the Poisson Model. The memoryless Poisson distribution is the predominant model used for analyzing traffic in traditional telephony networks. The Poisson process is characterized as a renewal process. In a Poisson process the inter-arrival times are exponentially distributed with a rate parameter λ: P{An ≤ t} = 1 – exp(-λt). The Poisson distribution is appropriate if the arrivals are from a large number of independent sources, referred to as Poisson sources. The distribution has a mean and variance equal to the parameter λ. The Poisson distribution can be visualized as a limiting form of the binomial distribution, and is also used widely in queuing models. There are a number of interesting mathematical properties exhibited by Poisson processes. Primarily, superposition of independent Poisson processes results in a new Poisson process whose rate is the sum of the rates of the independent Poisson processes. Further, the independent increment property renders a Poisson process memoryless. Poisson processes are common in traffic applications scenarios that consist of a large number of independent traffic streams. The reason behind the usage stems from Palm's Theorem which states that under suitable conditions, such large number of independent multiplexed streams approach a Poisson process as the number of processes grows, but the individual rates decrease in order to keep the aggregate rate constant. Traffic aggregation need not always result in a Poisson process. The two primary assumptions that the Poisson model makes are: [14] 1. The number of sources is infinite 2. The traffic arrival pattern is random.

In the compound Poisson model, the base Poisson model is extended to deliver batches of packets at once. The inter-batch arrival times are exponentially distributed, while the batch size is geometric. Mathematically, this model has two parameters, λ, the arrival rate, and ρ in (0,1), the batch parameter. Thus, the mean number of packets in a batch is 1/ ρ, while the mean inter-batch arrival time is 1/ λ. Mean packet arrivals over time period t are tλ/ ρ. The compound Poisson model shares some of the analytical benefits of the pure Poisson model: the model is still memoryless, aggregation of streams is still (compound) Poisson, and the steady-state equation is still reasonably simple to calculate, although varying batch parameters for differing flows would complicate the derivation. [14]

Markov models attempt to model the activities of a traffic source on a network, by a finite number of states. The accuracy of the model increases linearly with the number of states used in the model. However, the complexity of the model also increases proportionally with increasing number of states. An important aspect of the Markov model - the Markov Property, states that the next (future) state depends only on the current state. In other words, the probability of the next state, denoted by some random variable Xn+1, depends only on the current state, indicated by Xn, and not on any other state Xi, where i<n. The set of random variables referring to different states {Xn} is referred to as a Discrete Markov Chain.

Another attempt at providing a bursty traffic model is found in Jain and Routhier’s Packet Trains model. [17] This model was principally designed to recognize that address locality applies to routing decisions; that is, packets that arrive near each other in time are frequently going to the same destination. In generating a traffic model that allows for easier analysis of locality, the authors created the notion of packet trains, a sequence of packets from the same source, traveling to the same destination (with replies in the opposite direction). Packet trains are optionally sub-divided into tandem trailers. Traffic between a source and a destination usually consists of a series of messages back and forth. Thus, a series of packets go one direction, followed by one or more reply packets, followed by a new series in the initial direction. Traffic quantity is then a superposition of packet trains, which generates substantial bursty behavior. This refines the general conception of the compound Poisson model, which recognized that packets arrived in groups, by analyzing why they arrive in groups, and better characterizing the attributes of the group. Finally, the authors demonstrate that packet arrival times are not Poisson distributed, which led to a model that departs from variations on the Poisson theme. The packet train model is characterized by the following parameters and their associated probability distributions:

The train model is designed for analyzing and categorizing real traffic, not for generating synthetic loads for simulation. Thus, little claim has been made about the feasibility of packet trains for generating synthetic traffic. Given accurate parameters and distributions, generation should be straightforward, but derivation of these parameters is not addressed. [14]

Traffic models today

NS-2 is a popular network simulator; [18] PackMimeHTTP is a web traffic generator for NS-2, published in 2004. It does take long-range dependencies into account, and uses the Weibull distribution. Thus, it relies on heavy tails to emulate true self-similarity. Over most time scales, the effort is a success; only a long-running simulation would allow a distinction to be drawn. This follows suggestions from where it is suggested that self-similar processes can be represented as a superposition of many sources each individually modeled with a heavy-tailed distribution. It is clear that self-similar traffic models are in the mainstream. [14]

See also

Related Research Articles

The erlang is a dimensionless unit that is used in telephony as a measure of offered load or carried load on service-providing elements such as telephone circuits or telephone switching equipment. A single cord circuit has the capacity to be used for 60 minutes in one hour. Full utilization of that capacity, 60 minutes of traffic, constitutes 1 erlang.

<span class="mw-page-title-main">Exponential distribution</span> Probability distribution

In probability theory and statistics, the exponential distribution or negative exponential distribution is the probability distribution of the time between events in a Poisson point process, i.e., a process in which events occur continuously and independently at a constant average rate. It is a particular case of the gamma distribution. It is the continuous analogue of the geometric distribution, and it has the key property of being memoryless. In addition to being used for the analysis of Poisson point processes it is found in various other contexts.

<span class="mw-page-title-main">Erlang distribution</span> Family of continuous probability distributions

The Erlang distribution is a two-parameter family of continuous probability distributions with support . The two parameters are:

<span class="mw-page-title-main">Weibull distribution</span> Continuous probability distribution

In probability theory and statistics, the Weibull distribution is a continuous probability distribution. It models a broad range of random variables, largely in the nature of a time to failure or time between events. Examples are maximum one-day rainfalls and the time a user spends on a web page.

<span class="mw-page-title-main">Gamma distribution</span> Probability distribution

In probability theory and statistics, the gamma distribution is a two-parameter family of continuous probability distributions. The exponential distribution, Erlang distribution, and chi-squared distribution are special cases of the gamma distribution. There are two equivalent parameterizations in common use:

  1. With a shape parameter and a scale parameter .
  2. With a shape parameter and an inverse scale parameter , called a rate parameter.

A long-tailed or heavy-tailed distribution is one that assigns relatively high probabilities to regions far from the mean or median. A more formal mathematical definition is given below. In the context of teletraffic engineering a number of quantities of interest have been shown to have a long-tailed distribution. For example, if we consider the sizes of files transferred from a web server, then, to a good degree of accuracy, the distribution is heavy-tailed, that is, there are a large number of small files transferred but, crucially, the number of very large files transferred remains a major component of the volume downloaded.

In statistics, Poisson regression is a generalized linear model form of regression analysis used to model count data and contingency tables. Poisson regression assumes the response variable Y has a Poisson distribution, and assumes the logarithm of its expected value can be modeled by a linear combination of unknown parameters. A Poisson regression model is sometimes known as a log-linear model, especially when used to model contingency tables.

In queueing theory, a discipline within the mathematical theory of probability, a Markovian arrival process is a mathematical model for the time between job arrivals to a system. The simplest such process is a Poisson process where the time between each arrival is exponentially distributed.

Self-similar processes are types of stochastic processes that exhibit the phenomenon of self-similarity. A self-similar phenomenon behaves the same when viewed at different degrees of magnification, or different scales on a dimension. Self-similar processes can sometimes be described using heavy-tailed distributions, also known as long-tailed distributions. Examples of such processes include traffic processes, such as packet inter-arrival times and burst lengths. Self-similar processes can exhibit long-range dependency.

<span class="mw-page-title-main">Log-logistic distribution</span>

In probability and statistics, the log-logistic distribution is a continuous probability distribution for a non-negative random variable. It is used in survival analysis as a parametric model for events whose rate increases initially and decreases later, as, for example, mortality rate from cancer following diagnosis or treatment. It has also been used in hydrology to model stream flow and precipitation, in economics as a simple model of the distribution of wealth or income, and in networking to model the transmission times of data considering both the network and the software.

In queueing theory, a discipline within the mathematical theory of probability, Burke's theorem is a theorem asserting that, for the M/M/1 queue, M/M/c queue or M/M/∞ queue in the steady state with arrivals is a Poisson process with rate parameter λ:

  1. The departure process is a Poisson process with rate parameter λ.
  2. At time t the number of customers in the queue is independent of the departure process prior to time t.

This page lists articles related to probability theory. In particular, it lists many articles corresponding to specific probability distributions. Such articles are marked here by a code of the form (X:Y), which refers to number of random variables involved and the type of the distribution. For example (2:DC) indicates a distribution with two random variables, discrete or continuous. Other codes are just abbreviations for topics. The list of codes can be found in the table of contents.

<span class="mw-page-title-main">Relationships among probability distributions</span> Topic in probability theory and statistics

In probability theory and statistics, there are several relationships among probability distributions. These relations can be categorized in the following groups:

<span class="mw-page-title-main">Fork–join queue</span> Type of queue

In queueing theory, a discipline within the mathematical theory of probability, a fork–join queue is a queue where incoming jobs are split on arrival for service by numerous servers and joined before departure. The model is often used for parallel computations or systems where products need to be obtained simultaneously from different suppliers. The key quantity of interest in this model is usually the time taken to service a complete job. The model has been described as a "key model for the performance analysis of parallel and distributed systems." Few analytical results exist for fork–join queues, but various approximations are known.

<span class="mw-page-title-main">Poisson distribution</span> Discrete probability distribution

In probability theory and statistics, the Poisson distribution is a discrete probability distribution that expresses the probability of a given number of events occurring in a fixed interval of time or space if these events occur with a known constant mean rate and independently of the time since the last event. It is named after French mathematician Siméon Denis Poisson. The Poisson distribution can also be used for the number of events in other specified interval types such as distance, area, or volume. It plays an important role for discrete-stable distributions.

In queueing theory, a discipline within the mathematical theory of probability, a fluid queue is a mathematical model used to describe the fluid level in a reservoir subject to randomly determined periods of filling and emptying. The term dam theory was used in earlier literature for these models. The model has been used to approximate discrete models, model the spread of wildfires, in ruin theory and to model high speed data networks. The model applies the leaky bucket algorithm to a stochastic source.

In queueing theory, a discipline within the mathematical theory of probability, an M/D/1 queue represents the queue length in a system having a single server, where arrivals are determined by a Poisson process and job service times are fixed (deterministic). The model name is written in Kendall's notation. Agner Krarup Erlang first published on this model in 1909, starting the subject of queueing theory. An extension of this model with more than one server is the M/D/c queue.

In mathematics and telecommunications, stochastic geometry models of wireless networks refer to mathematical models based on stochastic geometry that are designed to represent aspects of wireless networks. The related research consists of analyzing these models with the aim of better understanding wireless communication networks in order to predict and control various network performance metrics. The models require using techniques from stochastic geometry and related fields including point processes, spatial statistics, geometric probability, percolation theory, as well as methods from more general mathematical disciplines such as geometry, probability theory, stochastic processes, queueing theory, information theory, and Fourier analysis.

In probability theory and statistics, the discrete Weibull distribution is the discrete variant of the Weibull distribution. It was first described by Nakagawa and Osaki in 1975.

References

  1. Willinger and Paxson (1998). "Where Mathematics Meets the Internet" (PDF). AMS.
  2. Park, Kihong; Willinger, Walter (2000). Self-similar network traffic and performance evaluation. New York: Wiley. doi:10.1002/047120644X.fmatter_indsub. ISBN   978-0-471-31974-0.
  3. Adas, A. (1997). "Traffic models in broadband networks". IEEE Communications Magazine. 35 (7): 82–89. CiteSeerX   10.1.1.23.1461 . doi:10.1109/35.601746. ISSN   0163-6804.
  4. Michiel, H.; Laevens, K. (1997). "Teletraffic engineering in a broad-band era". Proceedings of the IEEE. 85 (12): 2007–2033. doi:10.1109/5.650182. ISSN   0018-9219.
  5. Frost, V.S.; Melamed, B. (1994). "Traffic modeling for telecommunications networks". IEEE Communications Magazine. 32 (3): 70–81. doi:10.1109/35.267444. ISSN   0163-6804. S2CID   19019323.
  6. Chien-Hsing Wu; Huang-Pao Lin; Leu-Shing Lan (2002). "A new analytic framework for dynamic mobility management of PCS networks". IEEE Transactions on Mobile Computing. 99 (3): 208–220. doi:10.1109/TMC.2002.1081756.
  7. Thajchayapong, S.; Peha, J.M. (2006). "Mobility patterns in microcellular wireless networks" (PDF). IEEE Transactions on Mobile Computing. 5 (1): 52–63. doi:10.1109/tmc.2006.13. ISSN   1536-1233. S2CID   1175255.
  8. 1 2 Thomas M. Chen (2007). "Network Traffic Modeling". Southern Methodist University, Dallas, Texas.
  9. Kleinrock, Leonard (1975). Queueing systems (in German). New York: Wiley. ISBN   978-0-471-49110-1.
  10. Barakat, C.; Thiran, P.; Iannaccone, G.; Diot, C.; Owezarski, P. (2003). "Modeling internet backbone traffic at the flow level". IEEE Transactions on Signal Processing. 51 (8): 2111–2124. Bibcode:2003ITSP...51.2111B. CiteSeerX   10.1.1.2.8066 . doi:10.1109/tsp.2003.814521. ISSN   1053-587X.
  11. "Notes on effective bandwidths". Statistical Laboratory. Retrieved 2017-10-26.
  12. Nogueira, António; Salvador, Paulo; Valadas, Rui; Pacheco, António (2003). "Modeling Network Traffic with Multifractal Behavior" (PDF). Telecommunication Systems. 24 (2/4): 339–362. doi:10.1023/a:1026183318200. ISSN   1018-4864. S2CID   15332498. Archived from the original (PDF) on 2017-10-27.
  13. D.K. Arrowsmith, R.J. Mondragon (2006). "Modelling Network Data Traffic" (PDF). Queen Mary, University of London. Retrieved 2017-10-26.
  14. 1 2 3 4 5 6 7 8 Chandrasekaran, Balakrishnan (2006-11-27). "Survey of Network Traffic Models". Washington University in St. Louis - Computer Science & Engineering at WashU. Retrieved 2017-10-26.
  15. Wilson, Michael L. (2006). "A Historical View of Network Traffic Models". wustl.edu. Retrieved 2017-10-26.
  16. Jelenkovic, P.R.; Melamed, B. (1995). "Algorithmic modeling of TES processes" (PDF). IEEE Transactions on Automatic Control. 40 (7): 1305–1312. CiteSeerX   10.1.1.421.5082 . doi:10.1109/9.400470. ISSN   0018-9286.
  17. Jain, R.; Routhier, S. (1986). "Packet Trains--Measurements and a New Model for Computer Network Traffic". IEEE Journal on Selected Areas in Communications. 4 (6): 986–995. CiteSeerX   10.1.1.138.4617 . doi:10.1109/jsac.1986.1146410. ISSN   0733-8716.
  18. "nsnam". ns and nam Sourceforge home. 2010-04-05. Retrieved 2017-10-26.