Which of the following shows the number of units the market will buy in a given time period

Advanced Math and Statistics

Robert Kissell, Jim Poserina, in Optimal Sports Math, Statistics, and Fantasy, 2017

Poisson Distribution

The Poisson distribution is a discrete distribution that measures the probability of a given number of events happening in a specified time period. In finance, the Poisson distribution could be used to model the arrival of new buy or sell orders entered into the market or the expected arrival of orders at specified trading venues or dark pools. In these cases, the Poisson distribution is used to provide expectations surrounding confidence bounds around the expected order arrival rates. Poisson distributions are very useful for smart order routers and algorithmic trading.

Poisson Distribution Statistics1

Notation Poisson(λ)
Parameter λ>0
Distribution k=1,2,…,
Pdf λke−λk!
Cdf ∑i=1kλke−λ k!
Mean λ
Variance λ
Skewness λ−1/2
Kurtosis λ−1

Poisson Distribution Graph

Which of the following shows the number of units the market will buy in a given time period

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780128051634000049

Essential statistics for data analysis

Syed Naeem Ahmed, in Physics and Engineering of Radiation Detection (Second Edition), 2015

D.2 Poisson distribution

Poisson distribution represents the distribution of Poisson processes and is in fact a limiting case of the binomial distribution. By Poisson processes, we mean processes that are discrete, independent, and mutually exclusive.

The p.d.f. of a Poisson distribution is defined as

(9.3.31)f(x;μ)= μxe−μx!,

where x=0, 1, … represents the discrete random variable, such as ADC counts recorded by a detection system, and μ>0 is the mean. Figure 9.3.3 depicts this distribution for different values of μ. It is apparent that the width of the distribution increases with μ, which indicates that the uncertainty in measurement increases with an increase in the value of x.

Which of the following shows the number of units the market will buy in a given time period

Figure 9.3.3. Poisson probability density for different values of μ. The width of the distribution, which is a reflection of the uncertainty in measurements, increases with increase in μ.

Let us now apply the maximum likelihood method to determine the best estimate of the mean of a set of n measurements, assuming that the underlying process is Poisson in nature. The best way to do this is to use the maximum likelihood method we outlined earlier and applied in the previous section when discussing the binomial distribution. Since Poisson distribution is a discrete probability distribution, its likelihood function for a set of n measurements can be written as

(9.3.32)L(μ)=∏ i=1nf(xi,μ)=∏i=1n[μxi e−μxi!]=μ∑xie−nμ x1!x2!…xn!.

The log-likelihood function of L(μ) is

(9.3.33)l≡ln(L)=(∑i=1nxi)ln(μ)−nμ−ln(x1!x2!…xn!).

Following the maximum likelihood method (∂l/∂μ=0), we get

(9.3.34)∂∂μ[(∑ i=1nxi)ln(μ)−nμ−ln (x1!x2!…xn!)]=01μ*∑i=1 nxi−n=0μ*=1n∑i=1nxi.

This shows that the simple mean is the most probable value of a Poisson distributed variable. To determine the error in μ, we first take the second derivative of the log-likelihood function and then substitute it in Eq. (9.3.22):

(9.3.35)∂ 2l∂μ2=−1μ∑i=1n xiΔμ=[−∂2l∂μ2]−1/2=[−μ*2∑ i=1nxi]−1/2 =1n[∑i=1nxi]−1/2.

This is one of the most useful results of the Poisson distribution. It implies that if we make one measurement, the statistical error we should expect would simply be the square root of the measured quantity. For example, if we count the number of γ-ray photons coming from a radioactive source using a GM tube and get a number N, the statistical error we should expect will simply be N. Fortunately, most of the processes we encounter in the field of radiation detection and measurement, such as the activity of a radioisotope, photoelectric effect, and electron multiplication in a PMT tube, can all be very well described by Poisson statistics.

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780128013632000097

Probability Distributions

Ronald N. Forthofer, ... Mike Hernandez, in Biostatistics (Second Edition), 2007

5.2.1 Poisson Probabilities

The Poisson distribution arises from either of two models. In one model — quantities, for example — bacteria are assumed to be distributed at random in some medium with a uniform density of λ(lambda) per unit area. The number of bacteria colonies found in a sample area of size A follows the Poisson distribution with a parameter μ equal to the product of λ and A.

In terms of the model over time, we assume that the probability of one event in a short interval of length t1 is proportional to t1 — that is, Pr{exactly one event} is approximately λt1. Another assumption is that t1 is so short that the probability of more than one event during this interval is almost zero. We also assume that what happens in one time interval is independent of the happenings in another interval. Finally, we assume that λ is constant over time. Given these assumptions, the number of occurrences of the event in a time interval of length t follows the Poisson distribution with parameter μ, where μ is the product of λ and t.

The Poisson probability mass function is

Pr(X-x)=e-μμxx!forx=0,1,2……

where e is a constant approximately equal to 2.71828 and μ is the parameter of the Poisson distribution. Usually μ is unknown and we must estimate it from the sample data. Before considering an example, we shall demonstrate in Table 5.3 the use of the probability mass function for the Poisson distribution to calculate the probabilities when μ = 1 and μ = 2. These probabilities are not difficult to calculate, particularly when μ is an integer. There is also a recursive relation between the probability that X = x + 1 and the probability that X = x that simplifies the calculations:

Table 5.3. Calculation of poisson probabilities, Pr{X = x} = e−μ μx/x!, for μ = 1 and 2.

μ = 1μ = 2
xe−1 *1x/x! = Pr{X = x}e−1 *2x/x! = Pr{X = x}
0 0.3679 * 1/1 = 0.3679 0.1353 * 1/1 = 0.1353
1 0.3679 * 1/1 = 0.3679 0.1353 * 2/1 = 0.2707
2 0.3679 * 1/2 = 0.1839 0.1353 * 4/2 = 0.2707
3 0.3679 * 1/6 = 0.0613 0.1353 * 8/6 = 0.1804
4 0.3679 * 1/24 = 0.0153 0.1353 * 16/24 = 0.0902
5 0.3679 * 1/120 = 0.0031 0.1353 * 32/120 = 0.0361
6 0.3679 * 1/720 = 0.0005 0.1353 * 64/720 = 0.0120
7 0.3679 * 1/5040 = 0.0001 0.1353 * 128/5040 = 0.0034
8 0.1353 * 256/40320 = 0.0009
9 0.1353 * 512/362880 = 0.0002
1.0000 0.9999

Pr{X=x+1}=(μx+1 )Pr(X=x)

for x beginning at a value of 0. For example, for μ = 2,

Pr{X=3} =(2/3)Pr(X=2)=(2/3) 0.2707=0.1804

which is the value shown in Table 5.3.

These probabilities are also found in Appendix Table B3, which gives the Poisson probabilities for values of μ beginning at 0.2 and increasing in increments of 0.2 up to 2.0, then in increments of 0.5 up to 7, and in increments of 1 up to 17. Computer software can provide the Poisson probabilities for other values of μ (see Program Note 5.1 on the website). Note that the Poisson distribution is totally determined by specifying the value of its one parameter, μ. The plots in Figure 5.4 show the shape of the Poisson probability mass and cumulative distribution functions with μ = 2.

Which of the following shows the number of units the market will buy in a given time period

Figure 5.4. Poisson (μ = 2) probability mass and cumulative distribution functions.

The shape of the Poisson probability mass function with μ equal to 2 (the top plot in Figure 5.4) is similar to the binomial mass function for a sample of size 10 and π equal to 0.2 just shown. The cdf (the bottom plot in Figure 5.4) has the same general shape as that shown in the preceding binomial example, but the shape is easier to see here, since there are more values for the X variable shown on the horizontal axis.

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780123694928500108

Random Variables

Andrew F. Siegel, Michael R. Wagner, in Practical Business Statistics (Eighth Edition), 2022

The Poisson Distribution

The Poisson distribution, like the binomial, is a counted number of times something happens. The difference is that there is no specified number n of possible tries. Here is one way that it can arise. If an event happens independently and randomly over time and the mean rate of occurrence is constant over time, then the number of occurrences in a fixed amount of time will follow the Poisson distribution.13 The Poisson is a discrete distribution (because you can list the possibilities as 0, 1, 2, 3, …) and depends only on the mean number of occurrences expected.

Here are some random variables that might follow a Poisson distribution:

1.

The number of orders your firm receives tomorrow.

2.

The number of people who apply for a job tomorrow to your human resources division.

3.

The number of defects in a finished product.

4.

The number of calls your firm receives next week for help concerning an “easy-to-assemble” toy.

5.

A binomial number X when n is large and π is small.

The following figures show what the Poisson probabilities look like for a system expecting a mean of 0.5 occurrence (Fig. 7.5.1), two occurrences (Fig. 7.5.2), and 20 occurrences (Fig. 7.5.3). Note from the bell shape of Fig. 7.5.3 that the Poisson distribution is approximately normal when many occurrences are expected.

Which of the following shows the number of units the market will buy in a given time period

Fig. 7.5.1. The Poisson distribution with 0.5 occurrences expected is a skewed distribution. There is a high probability, 0.607, that no occurrences will happen at all.

Which of the following shows the number of units the market will buy in a given time period

Fig. 7.5.2. The Poisson distribution with two occurrences expected. The distribution is still somewhat skewed.

Which of the following shows the number of units the market will buy in a given time period

Fig. 7.5.3. The Poisson distribution with 20 occurrences expected. The distribution, although still discrete, is now fairly close to normal.

There are three important facts about a Poisson distribution. These facts, taken together, tell you how to find probabilities for a Poisson distribution when you know only its mean.

For a Poisson Distribution

1.

The standard deviation is always equal to the square root of the mean: σ=μ.

2.

The exact probability that a Poisson random variable X with mean μ is equal to a is given by the formula

P(X=a) =μaa!e−μ

where e = 2.71828… is a special number.14

3.

If the mean is large, then the Poisson distribution is approximately normal.

Example

How Many Warranty Returns?

Because your firm’s quality is so high, you expect only 1.3 of your products to be returned, on average, each day for warranty repairs. What are the chances that no products will be returned tomorrow? That one will be returned? How about two? How about three?

Because the mean (1.3) is so small, exact calculations are needed. Here are the details:

P(X=0)=1.3 00!e−1.3=11×0.27253=0.27253P(X=1)=1.311!e−1.3 =1.31×0.27253=0.35429P(X=2) =1.322!e−1.3=1.692×0.27253=0.23029P(X=3)=1.333! e−1.3=2.1976×0.27253=0.09979

From these basic probabilities, you could add up the appropriate probabilities for 0, 1, and 2 to also find the probability that two items or fewer will be returned. The probability is, then, 0.27253 + 0.35429 + 0.23029 = 0.857, or 85.7%.

To use Excel to compute these probabilities, you could use the function “=POISSON (value, mean, FALSE)” to find the probability that a Poisson random variable is exactly equal to some value, and you could use “=POISSON (value, mean, TRUE)” to find the probability that a Poisson random variable is less than or equal to the value. Here are the results:

Which of the following shows the number of units the market will buy in a given time period

Example

How Many Phone Calls?

Your firm handles 460 calls per day, on average. Assuming a Poisson distribution, find the probability that you will be overloaded tomorrow, with 500 or more calls received.

This may be computed directly in Excel using the formula

=1−POISSON( 500−1,460,TRUE)

to find the answer 0.0341, because the POISSON function finds the probability of being less than or equal to a given number; then we use the complement rule to find the probability of being greater; note that being “500 or more” is the complement of the event “499 or less.”

The mean, μ = 460, is given, and it follows that the standard deviation is σ=μ=460= 21.44761. The normal approximation is reasonable because the mean (460) is so large. Because the normal distribution is continuous, any value over 499.5 will round to 500 or more. For intuition, the standardized number of calls is

z=499.5−μσ=499.5−46021.44761=1.84

Our situation involves being more than 1.84 standard deviations above the mean, so we expect to see a few percent. Using the normal distribution, the answer is a probability of 1 − 0.967 = 0.033, which is fairly close to the more exact 0.341 found earlier. In conclusion, you may expect to be overloaded tomorrow with probability only about 3% (not very likely but within possibility).

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780128200254000075

Time Series Analysis: Methods and Applications

Konstantinos Fokianos, in Handbook of Statistics, 2012

2 Poisson regression modeling

The Poisson distribution is commonly used to model rate of random events that occur (arrive) in some fixed time interval. If we assume that λ denotes the rate of arrivals, then the distribution of the random variable Y, which denotes the number of arrivals in a fixed time interval, follows the Poisson distribution with probability mass function

(1)P[Y=y]=exp(−λ)λyy!,y=0,1,2 ,…

It is an elementary exercise to show that the mean and variance of Y are both equal to λ; E[Y] = Var[Y] = λ. In fact, this property characterizes the Poisson distribution. A related property is that the cumulant generating function of a Poisson random variable is given by KY(t)≡logMY(t)=λ( exp(t)−1), where MY (t) is the moment generating function of Y. This can be proved by simple calculations, but for this presentation, it is instructive to consider the Poisson distribution as a member of the natural exponential family of distributions.

Let f (x; θ) denote the density function of the natural exponential family with parameter θ, i.e.,

(2)f(x;θ)=h( x)exp(θx−b(θ)),x∈A,

where h(⋅), b(⋅) are known functions and A is a subset of R. Then, it is straightforward to show that the Poisson distribution is expressed as in (2) with θ = log λ, b(θ) = exp(θ), and h(x) = 1/x!. Using the fact that the cumulant generating function of (2) is equal to b(t + θ) − b(θ), the claim follows.

In most of the applications, count data are usually observed with some covariate information. For example, see the works of McCullagh and Nelder (1989, Section 6.3.2) where the authors study the relation between the type of ship, its year of construction, and its service period to the expected number of damage incidents using the logarithm of the aggregate months of service as an offset. (An offset is a continuous regression variable with corresponding known regression coefficient equal to 1.) In general, assume that X1, …, Xp are p regression variables observed jointly with a count response variable Y that follows the Poisson distribution. A possible regression model for association between the regressors and the expected value of Y given X1, …, Xp is

(3)λ=β0+∑i=1pβiXi.

This is an ordinary linear model with unknown regression coefficients βi, i = 0, …, p to be estimated. Model (3) poses several difficulties for fitting, because the parameter λ has to be positive. Nevertheless, in the context of time series and when the correlation among successive observations is positive, models such as (3) are quite useful; recall Fig. 1. A more natural choice for the regression modeling of count data is the so called log-linear model which is specified by

(4)logλ= β0+∑i=1pβiX i,

where the notation is as in (3). Regardless of the chosen model, a fact that remains true is that both (3) and (4) belong to the class of generalized linear models as introduced by Nelder and Wedderburn (1972) and elaborated further by McCullagh and Nelder (1989). Recall, that a generalized linear model consists of three components; the random component that belongs to the exponential family of distributions (2) with E[X] = μ, the systematic component η, and the link function g(⋅). The link function is a monotone twice differentiable function that is chosen by the user (or can be estimated). This function associates the random and systematic component via g(μ) = η. For the Poisson distribution, it is clear that both (3) and (4) introduce a generalized linear model with η=β0+∑i=1pβi Xi and g(λ) = λ (for (3)) and g(λ) = log λ (for (4)). Estimation and inference are based on the maximum likelihood theory – this topic has been described in several texts; see McCullagh and Nelder (1989) and Agresti (2002), for example. In the next section, we explore these ideas in the context of count time series.

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780444538581000120

Probability and statistics

Huw Fox, Bill Bolton, in Mathematics for Engineers and Technologists, 2002

Poisson distribution

The Poisson distribution for a variable λ is:

[23]P(k)=λke−λk!

for k = 0, 1, 2, 3, etc. The mean of this distribution is λ and the standard deviation is √λ. When the number n of trials is very large and the probability p small, e.g. n > 25 and p < 0.1, binomial probabilities are often approximated by the Poisson distribution. Thus, since the mean of the binomial distribution is np (equation [21]) and the standard deviation (equation [22]) approximates to √np when p is small, we can consider λ to represent np. Thus λ can be considered to represent the average number of successes per unit time or unit length or some other parameter.

Key point

Approximating the binomial distribution to the Poisson distribution

If p is the possibility of an event occurring and q the possibility that it does not occur, then q + p = 1 and so if we consider n samples, we have (q + p)n = 1 and the binomial expression gives:

(q +p)n=qn+nqn−1p+n(n−1)qn−2p22! +n(n−1)(n−2)qn−3 p33!+…

If p is small then q = 1 − p × 1 and with n large the first few terms have n − 1 approximating to n, n − 2 to n, etc. The binomial expression can thus be approximated to:

1n=1n+n1n−1p+nn1n−2p22!+nnn1n−3p3 3!+…1n=1+np+n2 p22!+n3p33!+…

If we let np = λ then:

1n=1+λ+λ22!+λ3 3!+…

But this is the series for e1 (see Table 7.1) and so 1n = e1. We can thus write the binomial expression as:

(q+p)n=e−λ[1+λ+λ22!+λ33!+…]

and so the terms in the expression are given by equation [23].

Example

2% of the output per month of a mass produced product have faults. What is the probability that of a sample of 400 taken that 5 will have faults?

Assuming the Poisson distribution, we have λ = np = 400 × 0.02 = 8 and so equation [23] gives for k = 5:

P(5)=85e−85!=0.093

Example

The output from a CNC machine is inspected by taking samples of 60 items. If the probability of a defective item is 0.0015, determine the probability of the sample having (a) two defective items, (b) more than two defective items.

(a)

We have n = 60 and p = 0.0015. Thus, assuming a Poisson distribution, we have λ = np = 60 × 0.0015 = 0.09 and so equation [23] gives the probability of two defective items as (0.092 × e−0.09)/2! = 3.7 × 10−3 or 0.37%.

(b)

The probability of there being more than two defective items is 1 − {P(0) + P(1) + P(2)] = 1 − e−λ{1 + λ + λ2/2!) = 1 − e−0.09{1 + 0.09 + 4.5 × 10−3} = 1.36 × 10−4 or 0.01%.

Example

There is a 1.5% probability that a machine will produce a faulty component. What is the probability that there will be at least 2 faulty items in a batch of 100?

Assuming the Poisson distribution can be used, we have λ = np = 100 × 0.015 = 1.5 and so the probability of at least 2 faulty items will be:

P(≥2)=1−P(0)−P(1)=1−1.500!e−1.5−1.511!e−1.5=0.442

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780750655446500102

Finding Probabilities of Error

ROBERT H. RIFFENBURGH, in Statistics in Medicine (Second Edition), 2006

POISSON EVENTS DESCRIBED

The Poisson distribution arises from situations in which there is a large number of opportunities for the event under scrutiny to occur but a small chance that it will occur on any one trial. The number of cases of bubonic plague would follow Poisson: A large number of patients can be found with chills, fever, tender enlarged lymph nodes, and restless confusion, but the chance of the syndrome being plague is extremely small for any randomly chosen patient. This distribution is named for Siméon Denis Poisson, who published the theory in 1837. The classic use of Poisson was in predicting the number of deaths of Prussian army officers from horse kicks from 1875 to 1894; there was a large number of kicks, but the chance of death from a randomly chosen kick was small.

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780120887705500538

Statistical Methods for Physical Science

Laurent Hodges, in Methods in Experimental Physics, 1994

2.2.3 The Poisson Distribution

The Poisson distribution is another common discrete distribution that is defined as follows. The Poisson probability distribution for x occurrences of a phenomenon when the average number of occurrences equals λ is

(2.5)p(x;λ)=e−λλxx!forx=0,1,2,…and someλ>0=0forothervalues ofx

The Poisson distribution occurs in several important situations. It is the limit of the binomial distribution b(x;n, p) when one lets n→∞ and p→0 in such a way that np remains fixed at the value λ>0. In practice, the Poisson distribution is an adequate approximation to the binomial distribution if n≥100 and p≤0.05.

The Poisson distribution also occurs in a more fundamental context, as a distribution associated with a “Poisson process,” which is a process satisfying the following conditions:

1.

The number of occurrences in an interval of time [t,t+Δt] is independent of the number of occurrences in any other interval that does not overlap with this interval.

2.

For small Δt, the probability of success in the time interval is approximately proportional to Δt.

3.

If the time interval Δt, is small enough, the probability of more than one occurrence in the interval is small compared to Δt.

For a Poisson process, the mean number of occurrences in an interval Δt can thus be expressed as γΔt, and 1/γ is the mean time between occurrences. (See Chapter 3 for more information on the Poisson process.)

An example of a Poisson process is the decay of radioactive nuclei in a sample containing a large number of these nuclei. Over 100 radioactive nuclei occur naturally (such as carbon-14, potassium-40, and uranium-238), plus many others are produced in laboratories. It has been observed, to a very high degree of accuracy, that every radioactive nucleus has a fixed probability of decaying in any particular time interval, and that this probability is independent of how long the nucleus has already been in existence. As an example, the probability that a uranium-238 nucleus will decay in 1 second is 4.87 × 10−18;; for one mole =6.022x1023 nuclei of uranium238, the mean number decaying in 1 second is thus 6.022 × 1023 × 4.87 × 10−18 = 2.93 × 106 nuclei.

The Poisson distribution has a particularly simple mean, E(X)=λ, and variance, V(X)=λ.

The Poisson cumulative distribution function does not have a simple form, though it can be easily calculated using a computer; it is

(2.6)P(X≤x)=∑v=0 rp(y;λ)forx=0,1,2…,n

Figure 5(a) shows the Poisson distributions for two values of λ, while Figure 5(b) shows the cumulative Poisson distributions for three values of λ.

Which of the following shows the number of units the market will buy in a given time period

Fig. 5. (a) The Poisson probability mass functions for λ = 2 and 15. (b) The cumulative Poisson distribution functions for λ = 2, 10, and 15.

Example 2.3:

Next, we consider radioactive decay. One curie (Ci) of a radioactive material is defined as the quantity of the material that leads to an average of 3.7 × 1010 decays per second. A more convenient unit in many circumstances is 1 picocurie (pCi)=10-12 Ci. What are the Poisson distributions for the number of decays of a 1-pCi sample in 1-sec and 1-min intervals?

The rate of decay of 1 pCi is 0.037/sec or 2.22/min. Setting λ=0.037 and λ=2.22 leads to the Poisson distributions shown in Table I. Note that there would be no decays in most 1-sec intervals, while in 1-min intervals the most common result would be two decays.

Table I. The Poisson Probability Mass Function and Cumulative Distribution Function for λ = 0.037 and λ = 2.22, Corresponding to the Radioactive Decaysper Second and per Minute for a 1-pCi Sample of a Radioactive Isotype

xλ = 0.037λ = 2.22
0 0.963676 0.963676 0.108609 0.108609
1 0.035656 0.999332 0.241112 0.349721
2 0.000660 0.999992 0.267635 0.617356
3 0.000008 1.000000 0.198050 0.815405
4 0.000000 1.000000 0.109918 0.925323
5 0.000000 1.000000 0.048803 0.974126
6 0.000000 1.000000 0.018057 0.992184
7 0.000000 1.000000 0.005727 0.997910
8 0.000000 1.000000 0.001589 0.999500
9 0.000000 1.000000 0.000392 0.999892
10 0.000000 1.000000 0.000087 0.999979

Example 2.4:

In a certain state with a stable population in which the annual number of suicides has averaged 300 per year over the past two decades, there are 330 suicides one year. The newspapers report that this is a 10% increase over the previous year, and a higher number than has occurred in the past 20 years, and speculate about the social or economic conditions that might be causing such a large increase. But is it so unusual? How often would the suicide rate be 330 or more? To answer this question, let us make the reasonable assumption that suicide is govemed by a Poisson process. A computer calculation shows that for λ = 300, P(X ≤ 329) = 0.954 = 1 − 0.046. That means that there is a 4.6% probability that, given an average rate of 300 suicides per year, the number of suicides in a given year is 330 or more. Apparently, the 330 total is not pmicularly unusual, and it would be advisable to watch the suicide rate for a few more years rather than speculate on the causes for the increase.

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/S0076695X08602525

Probability and statistics

Mary Attenborough, in Mathematics for Electrical Engineering and Computing, 2003

21.13 The Poisson distribution

The Poisson distribution is used to model processes where the distribution of the number of incidents occurring in any interval depends only on the length of that interval. Examples of such systems are:

1.

incoming telephone calls to an exchange during a busy period;

2.

customers arriving at a checkout counter in a supermarket from 4 to 6p.m.;

3.

accidents on a busy stretch of the M1;

4.

number of misprints in a book.

When modelling situations in a Poisson process we use four assumptions:

1.

If A is the event of n incidents in an interval and B the event of m incidents in another non-overlapping interval then A and B are independent, that is, p(A∩B) = p(A)p(B).

2.

If A is the event of n incidents in an interval then P(A) depends only on the length of the interval– not on the starting point of the interval.

3.

The probability of exactly one incident in a small interval is approximately proportional to the length of that interval, that is, P1(t)≈λtfor small t.

4.

The probability of more than one incident in a small interval is negligible. Thus, for small t, P2 (t) ≈0 and we can also say that

It follows that P0(t) + P1(t) ≈ 1 and as by assumption (3), P1(t) ≈ λt we get P0(t) ≈ 1 – λt. We now think about the number of incidents in an interval of time of any given length, (0, t), where t is no longer small. We can divide the interval into pieces of length h, where h is small, and use the assumptions above. We can see that in each small interval of length h there is either no event or a single event. Therefore, each small interval is approximately behaving like a Bernoulli trial. This means that we can approximate the events in the interval (0, t) by using the Binomial distribution for the number of successes r in n trials. The probability of r incidents in n intervals, where the probability of an incident in any one interval is λh, is given by

Pr(t)=limh→0(nr)(λh)r(1−λh)n−r.

Substituting h = t / n gives

Pr(t)=limn→∞n!(n−r)!r! λrtrnr(1−λtn) n−r.

We can reorganize this expression, by taking out of the limit terms not involving n

Pr(t)= (λt)rr!limn→∞n!(n−r)!nr(1−λt n)n(1−λtn)−r.

We can rewrite the first term inside the limit to give

Pr(t)=(λt)r r!limn→∞nnn−1n⋯n−rn(1−λtn)n(1−λtn)−r.

Now we notice that the first term inside the limit is made up of the product of r fractional expressions, which each have a term in n on the top and bottom lines. These will all tend to 1 as n tends to ∞. The last term is similar to the limit that we saw in Chapter 7 when calculating the value of e. There, we showed that

limn→∞(1+1n)n=e

and by a similar argument we could show that

limn→∞(1+xn)n=ex.

It, therefore follows that

limn→∞(1+λtn)n=e −λt.

The last expression involves a negative power of (1− λt/n), which will tend to 1 as n tends to ∞.

This gives the Poisson distribution as

Pr(t)=(λt)rr!e−λt.

This is an expression in both r and t where r is the number of events and t is the length of the time interval being considered. We usually consider the probability of r events in an interval of unit time, which gives the Poisson distribution as

Pr=λrr!e−λ

where λ is the expected number of incidents in unit time.

The mean and variance of the Poisson distribution

The Poisson distribution was introduced by considering the probability of a single event in a small interval of length h as (λh). We then used the binomial distribution, with θ = λh and h = t/n and n tending to ∞, to derive the expression for the Poisson distribution. As the mean of the binomial distribution is nθ it would make sense that the mean of the Poisson distribution is nλh. Using n = t/h we get the mean as λt over an interval of length t and therefore the mean is λ over an interval of unit length.

By a similar argument we know that the variance of the Binomial distribution is nθ(1 – θ). Substituting θ = λh we get the variance as nλh(1 – λh). As n tends to infinity and h to 0 we get the limit λt. Therefore, the variance in unit time is λ.

Example 21.16

The average number of ‘page not found’ errors on a web server is 36 in a 24-h period. Find the probability in a 60-min period that:

(a)

there are no errors;

(b)

there is exactly one error;

(c)

There are at most two errors;

(d)

there are more than three errors.

Solution Assuming that the above process is a Poisson process, then we have that the average number of errors in 24 h is 36 and therefore the average is 1.5 in 1 h. As the mean is λ, we can now assume that the number of errors in 1 h follows a Poisson distribution with λ = 1.5, giving

Pr=(1.5)rr!e−1 .5.

(a)

We want to find P(no errors) = Po = ((1.5)0/0!)e−1.5= e−1.5 ≈0.2231.

(b)

We want to find P(exactly one error) = P1= (1.5)1 /1!e−1.5 = 1.5e−1.5 ≈0.3347.

(c)

p (at most two errors) = Po + P1 + P2= 0.2231 + 0.3347 + (1.5)2/2!e−1.5 = 0.8088.

(d)

P(more than three errors) = 1 –P (at most three errors) = 1 – (P0+ P1+ P2+ P3).

Using the result from Part (c) we get

P(morethanthreeerrors)=1−0.8808−(1.5)3/3!e−1.5 ≈1−0.8808−0.1255=0.0657.

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780750658553500474

Special Types of Regression

Donna L. Mohr, ... Rudolf J. Freund, in Statistical Methods (Fourth Edition), 2022

13.3 Poisson Regression

The Poisson distribution is widely used as a model for count data. As discussed in Section 2.3, it is frequently appropriate when the counts are of events in specific regions of time or space. Dependent variables that might be modeled using the Poisson regression would include the

number of fatal auto accidents during a year at intersections as a function of lane width,

number of service interruptions during a month for a network server as a function of usage, and

number of fire ant colonies in an acre of land as a function of tree density.

There is no fixed upper limit on the possible number of events. Recalling the properties of the Poisson, there is a single parameter, μ , which is the expected number of events. It is essential that μ be positive, and the regression function must enforce this.

Case Study 13.1

Warner (2007) studied the proportions of robberies where a gun was used as a function of the characteristics of the neighborhoods where the robberies occurred. Using logistic regression, the author related the probability the robbery would involve a gun to several independent variables:

Disadvantaged, a blend of poverty level, % female-headed households, and other economic variables, with higher scores indicating poorer neighborhoods

Percent population that are young black males

Faith in police, a score obtained by surveying residents of the neighborhood, with high scores indicating a greater trust of police

Perceived oppositional values, a score obtained by surveying residents of the neighborhood, with high scores indicating more opposition to mainstream attitudes toward drugs and crime

Table 13.5 shows the results of the bivariate (dependent variable against a single independent variable) logistic regressions. Also shown are the means (M) and standard deviations (SD) for the independent variables.

Table 13.5. Results of logistic regressions for gun use in robberies.

Ind. VariableM (SD)β^(std. error)zOdds Ratio
Disadvantaged 0.00 (1.00) 0.33 (0.15) 2.16 1.39
% young black male 3.26 (2.11) 0.14 (0.08) 1.75 1.14
Faith in police 3.40 (0.17) −3.37(1.13) −2.97 0.03
Perceived opp. values 40.70 (7.30) −0.02(0.02) −1.15 0.98

Disadvantaged shows a significant positive association; that is, the more disadvantaged a neighborhood, the greater the probability a robbery will involve a gun. Faith in police shows a significant negative association; that is, the greater the neighborhood’s faith in police, the less the probability the robbery will involve a gun. Neither of the other independent variables showed a significant association.

The information regarding the standard deviations of the independent variable is important in helping understand each variable’s impact. Disadvantaged has SD=1.0, so that a moderately low score might be 1 point below the mean and a high score might be 1 point above the mean. Moving from a moderately low score to a moderately high score on Disadvantaged would shift the odds ratio by exp(2×1×0.33)=1.93. By contrast, Faith in police, which has a much larger coefficient, only has SD= 0.17. Moving from a moderately high score to a moderately low score on Faith in police would shift the odds by exp(2×.17×3.37)=3.14. Thus, the difference in impact of the two variables is not as large as we would think if we only examined the coefficients. Note that we have described a shift from high to low for Faith in police, which changed the sign on the coefficient. This made the comparison of the two independent variables easier.

Poisson regression assumes each yi follows a Poisson distribution with mean μi, where

ln(μi)=β0+β1x1+β2x2+⋯+βmxm.

The linear expression may take on either positive or negative values, but μi=exp(β0+β1x1+β2x2+⋯+βmx m) will always be positive. Note that the link function is the logarithmic function. The proper method of fitting this model is via ML. Likelihood ratio tests replace F tests, and Wald χ2 tests will replace t tests.

Example 13.3

Hypoglycemic Incidents

A hospital tracks the number of hypoglycemic incidents among diabetic patients recovering from cardiovascular surgery. The dependent variable is the number of incidents experienced by a patient during the first 72 hours post-surgery. The research question is whether a patient’s age can be related to the frequency of incidents. An artificial data set illustrating this situation is given in Table 13.6.

Table 13.6. Number of incidents of hypoglycemia.

OBSAGEHYPOGOBSAGEHYPOG
1 52 0 16 62 2
2 74 1 17 66 1
3 57 1 18 71 0
4 73 0 19 50 1
5 72 1 20 64 2
6 53 0 21 65 0
7 72 2 22 67 0
8 75 0 23 75 1
9 57 0 24 57 0
10 69 0 25 56 1
11 63 1 26 55 0
12 73 1 27 70 0
13 67 1 28 67 0
14 64 3 29 66 1
15 76 0 30 74 0

Solution

We will model each person’s number of incidents as having a Poisson distribution where the expected number (μ) is a function of AGE,

ln(μ)=β0+β1age.

The SAS System’s PROC GENMOD was used to fit this model, and a portion of the output is shown in Table 13.7. Several portions of this printout deserve comment.

Table 13.7. Poisson regression results for hypoglycemia.

The GENMOD Procedure
Criteria For Assessing Goodness of Fit
CriteriondfValueValue/df
Deviance 28 31.1272 1.1117 (*1*)
Scaled Deviance 28 31.1272 1.1117
Pearson Chi-Square 28 27.9940 0.9998
Scaled Pearson X2 28 27.9940 0.9998
Log Likelihood −28.1089 (*2*)
Algorithm converged.
Analysis Of Parameter Estimates
ParameterdfEstimateStandard ErrorWald 95% Confidence LimitsChi-SquarePr &gt; ChiSq
Intercept 1 −0.3486 1.9394 −4.1496 3.4525 0.03 0.8574
age 1 −0.0009 0.0295 −0.0586 0.0569 0.00 0.9764 (*3*)
Scale 0 1.0000 0.0000 1.0000 1.0000

Note: The scale parameter was held fixed.

(*1*) Deviance is a measure of lack of fit for the proposed model versus a saturated model that essentially includes one parameter for every observation. The saturated model represents a kind of gold standard. Comparing the Deviance (31.1272) to the critical values for a chi-squared distribution with degrees of freedom as shown on the printout (28) gives a very rough lack of fit test. That is, large values of deviance indicate the model is not a good fit. As a rough rule-of-thumb, we expect deviance divided by its degrees of freedom to be in the vicinity of 1. The value here is 1.1117, indicating that the model fits the data reasonably well.

(*2*) ln(L) of the current model ( −28.1089), which is useful when constructing customized likelihood ratio tests comparing full and reduced models.

(*3*) The χ2 tests for each individual independent variable, analogous to the individual t tests in ordinary regression. For AGE, X2=0.00, with p value=0.9764.

There is no significant evidence that age is related to the frequency of hypoglycemic incidents.

Sometimes each observation is a count from a region that varies greatly in size. For example, we might have y=number of flaws in a Mylar sheet, but some sheets are quite large and others are small. In this situation, size is an important part of the expected count. The independent variables are assumed to influence the rate per unit of size, denoted λ. The rate must be positive. Given a set of independent variables x1,x2,…,xm , we model

ln(λ)=β0+β1x1+β2x2+⋯+βmx m.

If observation yi comes from an observational unit with size si, then y i has the Poisson distribution with expected value μi=λisi and

ln(μi)=β0+ β1x1+β2x2+⋯+βmxm+ln(si).

At first glance, the term ln(si) may seem like just another independent variable in the Poisson regression. However, its coefficient is identically 1, so that no parameter need be estimated for it. This is called an offset variable, and all Poisson regression software will allow you to indicate such a size marker. Sometimes size is only specified up to a constant of proportionality. That is, we might not know exactly the size of units i and i′, but we know that unit i is twice the size of unit i′. This suffices, as the unknown proportionality constant will become an additive constant once logarithms are computed, and be combined with the intercept β0.

Example 13.4

Occupational Fatality Rates

Bailer et al. (1997) published an article showing how Poisson regression could be an important tool in safety research. Table 13.8 shows their counts of fatalities in the agriculture, forestry, and fishing industries and estimates of the number of workers in those industries. Figure 13.4 graphs the rates per 1,000 workers (number of fatalities×1,000/number of workers). We would like to see that fatality rates are declining, but is there any evidence that this is so?

Table 13.8. Fatalities and number of workers.

YearFatalitiesWorkersYearFatalitiesWorkers
1983 511 28,50,803 1988 506 26,49,044
1984 530 27,67,829 1989 491 26,65,645
1985 566 26,67,323 1990 464 26,14,612
1986 499 26,79,587 1991 484 26,66,477
1987 529 27,09,966 1992 468 25,81,603

Which of the following shows the number of units the market will buy in a given time period

Figure 13.4. Fatality Rates Among Workers for Example 13.4.

Solution

We will model the number of fatalities each year as a Poisson variable with mean μi=λis i, where λi is the rate of fatalities per worker in year i, and si is the number of workers in these industries during year i. To model a trend in time, we use

ln(λi)=β 0+β1i,

where i=yea r−1982. The link function is the logarithmic function, and the offset variable is ln(si). The SAS System’s PROC GENMOD yielded β^1=−0.0073 with a standard error of 0.0049 and Wald’s χ2=2.21,pvalue=0.1373. Hence, there is no significant evidence of a linear trend in the fatality rate over this time period. In a second analysis where fatalities and number of workers were subdivided by gender and age, the authors found that rates were decreasing significantly among male workers, but increasing among female workers.

13.3.1 Choosing Between Logistic and Poisson Regression

When data are presented as results for individual observations, as in Example 13.2 and Example 13.3, the choice between logistic regression and Poisson regression is usually clear. In Example 13.2, the dependent variable was whether or not a vaccine recipient developed a fever, which happened to be coded as 0s and 1s but could have been Y/N or any other abbreviation. At the individual level, the dichotomous variable is whether or not a success has occurred. This is the type of dependent variable where logistic regression is helpful as we attempt to model the probability of a success.

In Example 13.3, the dependent variable is truly quantitative, the number of hypoglycemic incidents experienced by a patient. This number happens to almost always be 0 or 1, but is not necessarily one of these two values. In fact, the data contains two individuals who had more than one incident, though another sample might not have had any. This is the type of situation where Poisson regression is helpful as we attempt to model the expected number of incidents per patient.

The choice is somewhat less distinct when the data have been aggregated for groups of similar individuals, as in Example 13.1 and Example 13.4. In Example 13.4, we treated the dependent variable as yi=number of fatalities in year i, assumed to have a Poisson distribution with number of workers as an offset variable. However, we could also treat yi as a binomial random variable with ni= number of workers. After all, a worker cannot have more than one fatality! In fact, these two approaches would give very similar fitted values for μy |x=px, because the Poisson and binomial are very similar when n is large and p is very small.

By contrast, Example 13.1 can only use logistic regression. First, at the individual level, our data is whether or not a mouse developed a tumor. This is a binary dependent variable. If, at every concentration, the probability of a tumor stayed small, we could still use Poisson regression if it were more convenient. Our dependent variable would be the number of mice with tumors within each sample at a given concentration. However, for this data set, p ranges from small to large. The approximation of the binomial via the Poisson deteriorates. Moreover, the link function for the logistic regression will keep all the fitted values for the probability between 0 and 1. The link function for Poisson regression will keep them greater than 0, but is likely to return some greater than 1.

As best we can, the choice between logistic regression and Poisson regression should match the nature of the dependent variable at the level of the individual observation. In certain cases, however, where the proportion of successes out of the total number of trials is quite small, we may analyze the data either way. Be aware, however, that the regression coefficients are giving different information. For Poisson regression, they reflect the influence of an independent variable on the ln(p), but for logistic regression they reflect the influence on ln(odds).

Case Study 13.2

Darby et al. (2009) studied auto collision records for over 16,000 employees of a large British telecommunications firm. Each of these employees was the driver of a company car or van, and the dependent variable in question was each person’s number of collisions, in a company vehicle, during the past three years. In addition to the more traditional risk factors of gender and age group, the authors attempted to assess whether certain personality traits were associated with a change in the rate of accidents. For data on personality traits, they had each employee’s answers on a questionnaire given to them at the time they were approved to drive a company vehicle.

Since the data is in the form of counts (many of them zeroes), the authors chose Poisson regression as the primary means of analysis. Since some workers drove very little during the week and others a great deal, ln(# hours driven per week) was used as an offset variable. With this sample size, the authors were able to fit a model with a large number of independent variables. We cite a few of their results.

Dummy variables were used to code different age categories, with the 50+ age category acting as the reference group. For ages 21 to 25, β^=0.366 with p value<0.001. To interpret this, consider two persons with all independent variables equal except that one is in the 21 to 25 age category and the other is in the 50+ age category. Their fitted accident rates will differ by ln(λ^21−25)−ln(λ^50+)=0.366. Hence, λ^21− 25∕λ^50+=exp(0.366)=1.44. That is, the fitted rate in the 21 to 25 age category is 44% higher than that in the 50+ age category, all other variables being equal.

Persons scoring high on the aggressive/impulsive personality trait had a substantially higher rate of accidents (β^=0.529,p value<0.001). The structured personality trait had no significant relationship with accident rate (β^=−0.14,p value=0.823).

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780128230435000138

Which of the following is an example of a pricing strategy that focuses on customers needs?

Value-based pricing. In value-based pricing, the perceived value to the customer is primarily based on how well it's suited to the needs and wants of each customer.

What is cost

What is cost-based pricing? Cost-based pricing is a pricing method that is based on the cost of production, manufacturing, and distribution of a product. Essentially, the price of a product is determined by adding a percentage of the manufacturing costs to the selling price to make a profit.

What is good value pricing in marketing?

A good value pricing strategy focuses on features, not value. The goal is to make consumers believe they are getting a good product at a fair price. When creating marketing campaigns for these types of products, marketers don't need to focus on building a lot of additional value.

What is value

Value-added pricing refers to the practice of pricing a product based on the perceived value that products and their features add for a customer. Sellers attempt to determine what customers believe the value of a particular feature of the product is worth and price the product accordingly.