Skip to main content
Statistics LibreTexts

5.39: Benford's Law

  • Page ID
    10472
  • \( \newcommand{\vecs}[1]{\overset { \scriptstyle \rightharpoonup} {\mathbf{#1}} } \) \( \newcommand{\vecd}[1]{\overset{-\!-\!\rightharpoonup}{\vphantom{a}\smash {#1}}} \)\(\newcommand{\id}{\mathrm{id}}\) \( \newcommand{\Span}{\mathrm{span}}\) \( \newcommand{\kernel}{\mathrm{null}\,}\) \( \newcommand{\range}{\mathrm{range}\,}\) \( \newcommand{\RealPart}{\mathrm{Re}}\) \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\) \( \newcommand{\Argument}{\mathrm{Arg}}\) \( \newcommand{\norm}[1]{\| #1 \|}\) \( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\) \( \newcommand{\Span}{\mathrm{span}}\) \(\newcommand{\id}{\mathrm{id}}\) \( \newcommand{\Span}{\mathrm{span}}\) \( \newcommand{\kernel}{\mathrm{null}\,}\) \( \newcommand{\range}{\mathrm{range}\,}\) \( \newcommand{\RealPart}{\mathrm{Re}}\) \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\) \( \newcommand{\Argument}{\mathrm{Arg}}\) \( \newcommand{\norm}[1]{\| #1 \|}\) \( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\) \( \newcommand{\Span}{\mathrm{span}}\)\(\newcommand{\AA}{\unicode[.8,0]{x212B}}\)

    \(\newcommand{\R}{\mathbb{R}}\) \(\newcommand{\N}{\mathbb{N}}\) \(\newcommand{\Z}{\mathbb{Z}}\) \(\newcommand{\E}{\mathbb{E}}\) \(\newcommand{\P}{\mathbb{P}}\) \(\newcommand{\var}{\text{var}}\) \(\newcommand{\sd}{\text{sd}}\)

    Benford's law refers to probability distributions that seem to govern the significant digits in real data sets. The law is named for the American physicist and engineer Frank Benford, although the law was actually discovered earlier by the astronomer and mathematician Simon Newcomb.

    To understand Benford's law, we need some preliminaries. Recall that a positive real number \(x\) can be written uniquely in the form \(x = y \cdot 10^n\) (sometimes called scientific notation) where \(y \in \left[\frac{1}{10}, 1\right)\) is the mantissa and \(n \in \Z\) is the exponent (both of these terms are base 10, of course). Note that \[ \log x = \log y + n \] where the logarithm function is the base 10 common logarithm instead of the usual base \(e\) natural logarithm. In the old days BC (before calculators), one would compute the logarithm of a number by looking up the logarithm of the mantissa in a table of logarithms, and then adding the exponent. Of course, these remarks apply to any base \(b \gt 1\), not just base 10. Just replace 10 with \(b\) and the common logarithm with the base \(b\) logarithm.

    Distribution of the Mantissa

    Distribution Functions

    Suppose now that \(X\) is a number selected at random from a certain data set of positive numbers. Based on empirical evidence from a number of different types of data, Newcomb, and later Benford, noticed that the mantissa \(Y\) of \(X\) seemed to have distribution function \(F(y) = 1 + \log y\) for \(y \in [1/10, 1)\). We will generalize this to an arbitrary base \(b \gt 1\).

    The Benford mantissa distribution with base \( b \in (1, \infty) \), is a continuous distribution on \([1/b, 1)\) with distribution function \( F \) given by \[ F(y) = 1 + \log_b y, \quad y \in [1/b, 1) \] The special case \(b = 10\) gives the standard Benford mantissa distribution.

    Proof

    Note that \( F \) is continuous and strictly increasing on \( [1 / b, 1) \) with \(F(1 / b) = 0 \) and \( F(1) = 1 \).

    The probability density function \( f \) is given by \[ f(y) = \frac{1}{y \ln b}, \quad y \in [1/b, 1)\]

    1. \( f \) is decreasing with mode \( y = \frac{1}{b} \).
    2. \( f \) is concave upward.
    Proof

    These results follow from the CDF \(F\) above and standard calculus. Recall that \( f = F^\prime \).

    Open the Special Distribution Simulator and select the Benford mantissa distribution. Vary the base \( b \) and note the shape of the probability density function. For various values of \( b \), run the simulation 1000 times and compare the empirical density function to the probability density function.

    The quantile function \( F^{-1} \) is given by \[ F^{-1}(p) = \frac{1}{b^{1-p}}, \quad p \in [0, 1] \]

    1. The first quartile is \( F^{-1}\left(\frac{1}{4}\right) = \frac{1}{b^{3/4}} \)
    2. The median is \( F^{-1}\left(\frac{1}{2}\right) = \frac{1}{\sqrt{b}} \)
    3. The third quartile is \( F^{-1}\left(\frac{3}{4}\right) = \frac{1}{b^{1/4}} \)
    Proof

    The formula for \( F^{-1}(p) \) follows by solving \( F(x) = p \) for \( x \) in terms of \( p \).

    Numerical values of the quartiles for the standard (base 10) distribution are given in an exercise below.

    Open the special distribution calculator and select the Benford mantissa distribution. Vary the base and note the shape and location of the distribution and probability density functions. For selected values of the base, compute the median and the first and third quartiles.

    Moments

    Assume that \( Y \) has the Benford mantissa distribution with base \( b \in (1, \infty) \).

    The moments of \( Y \) are \[ \E\left(Y^n\right) = \frac{b^n - 1}{n b^n \ln b}, \quad n \in (0, \infty) \]

    Proof

    For \( n \gt 0 \), \[ \E\left(Y^n\right) = \int_{1/b}^1 y^n \frac{1}{y \ln b} dy = \frac{1}{\ln b} \int_{1/b}^1 y^{n-1} dy = \frac{1 - 1 \big/ b^n}{n \ln(b)} \]

    Note that for fixed \( n \gt 0 \), \( \E(Y^n) \to 1 \) as \( b \downarrow 1 \) and \( \E(Y^n) \to 0 \) as \( b \to \infty \). We will learn more about the limiting distribution below. The mean and variance follow easily from the general moment result.

    Mean and variance

    1. The mean of \(Y\) is \[\E(Y) = \frac{b - 1}{b \ln b}\]
    2. the variance of \(Y\) is \[\var(Y) = \frac{b - 1}{b^2 \ln b} \left[ \frac{b + 1}{2} - \frac{b - 1}{\ln b} \right] \]

    Numerical values of the mean and variance for the standard (base 10) distribution are given in an exercise below.

    In the Special Distribution Simulator, select the Benford mantissa distribution. Vary the base \( b \) and note the size and location of the mean \( \pm \) standard deviation bar. For selected values of \( b \), run the simulation 1000 times and compare the empirical mean and standard deviation to the distribution mean and standard deviation.

    Related Distributions

    The Benford mantissa distribution has the usual connections to the standard uniform distribution by means of the distribution function and quantile function given above.

    Suppose that \( b \in (1, \infty) \).

    1. If \( U \) has the standard uniform distribution then \( Y = b^{-U} \) has the Benford mantissa distribution with base \( b \).
    2. If \( Y \) has the Benford mantissa distribution with base \( b \) then \( U = -\log_b Y \) has the standard uniform distribution.
    Proof
    1. If \( U \) has the standard uniform distribution then so does \( 1 - U \) and hence \( Y = F^{-1}(1 - U) = b^{-U}\) has the Benford mantissa distribution with base \( b \).
    2. The CDF \( F \) is strictly increasing on \( [b^{-1}, 1) \). Hence if \( Y \) has the Benford mantissa distribution with base \( b \) then \( F(Y) = 1 + \log_b(Y) \) has the standard uniform distribution and hence so does \( 1 - F(Y) = -\log_b(Y) \).

    Since the quantile function has a simple closed form, the Benford mantissa distribution can be simulated using the random quantile method.

    Open the random quantile experiment and select the Benford mantissa distribution. Vary the base \( b \) and note again the shape and location of the distribution and probability density functions. For selected values of \( b \), run the simulation 1000 times and compare the empirical density function, mean, and standard deviation to their distributional counterparts.

    Also of interest, of course, are the limiting distributions of \( Y \) with respect to the base \( b \).

    The Benford mantissa distribution with base \( b \in (1, \infty) \) converges to

    1. Point mass at 1 as \( b \downarrow 1 \).
    2. Point mass at 0 as \( b \uparrow \infty \).
    Proof

    Note that the CDF of \(Y\) above can be written as \( F(y) = 1 + \ln(y) \big/ \ln(b) \) for \( 1/b \le y \lt 1 \), and of course we also have \( F(y) = 0 \) for \( y \lt 1/b \) and \( F(y) = 1 \) for \( y \ge 1 \).

    1. As \( b \downarrow 1 \), \( 1/b \uparrow 1 \), and \( 1 + \ln(y) \big/ \ln(b) \to 1 \), so in the limit we have \( F(y) = 0 \) for \( y \lt 1 \) and \( F(y) = 1 \) for \( y \gt 1 \).
    2. As \( b \uparrow \infty \), \( 1/b \downarrow 0 \), and again \( 1 + \ln(y) \big/ \ln(b) \to 1 \), so in the limit we have \( F(y) = 0\) for \( y \lt 0 \) and \( F(y) = 1 \) for \( y \gt 0). \)

    Since the probability density function is bounded on a bounded support interval, the Benford mantissa distribution can also be simulated via the rejection method.

    Open the rejection method experiment and select the Benford mantissa distribution. Vary the base \( b \) and note again the shape and location of the probability density functions. For selected values of \( b \), run the simulation 1000 times and compare the empirical density function, mean, and standard deviation to their distributional counterparts.

    Distributions of the Digits

    Assume now that the base is a positive integer \(b \in \{2, 3, \ldots\}\), which of course is the case in standard number systems. Suppose that the sequence of digits of our mantissa \(Y\) (in base \(b\)) is \(\left(N_1, N_2, \ldots\right)\), so that \[ Y = \sum_{k=1}^\infty \frac{N_k}{b^k} \] Thus, our leading digit \(N_1\) takes values in \(\{1, 2, \ldots, b - 1\}\), while each of the other significant digits takes values in \(\{0, 1, \ldots, b - 1\}\). Note that \( \left(N_1, N_2, \ldots\right) \) is a stochastic process so at least we would like to know the finite dimensional distributions. That is, we would like to know the joint probability density function of the first \(k\) digits for every \( k \in \N_+ \). But let's start, appropriately enough, with the first digit law. The leading digit is the most important one, and fortunately also the easiest to analyze mathematically.

    First Digit Law

    \( N_1 \) has probability density function \( g_1 \) given by \(g_1(n) = \log_b \left(1 + \frac{1}{n} \right) = \log_b(n + 1) - \log_b(n)\) for \(n \in \{1, 2, \ldots, b - 1\}\). The density function \( g_1 \) is decreasing and hence the mode is \( n = 1 \).

    Proof

    Note that \(N_1 = n\) if and only if \(\frac{n}{b} \le Y \lt \frac{n + 1}{b}\) for \( n \in \{1, 2, \ldots, b - 1\} \). Hence using the PDF of \(Y\) above, \[ \P(N_1 = n) = \int_{n/b}^{(n+1)/b} \frac{1}{y \ln b} dy = \log_b\left(\frac{n+1}{b}\right) - \log_b\left(\frac{n}{b}\right) = \log_b(n + 1) - \log_b(n) \]

    Note that when \( b = 2 \), \( N_1 = 1 \) deterministically, which of course has to be the case. The first significant digit of a number in base 2 must be 1. Numerical values of \( g_1 \) for the standard (base 10) distribution are given in an exercise below.

    In the Special Distribution Simulator, select the Benford first digit distribution. Vary the base \( b \) with the input control and note the shape of the probability density function. For various values of \( b \), run the simulation 1000 times and compare the empirical density function to the probability density function.

    \( N_1 \) has distribution function \( G_1 \) given by \( G_1(x) = \log_b \left(\lfloor x \rfloor + 1\right) \) for \( x \in [1, b - 1] \).

    Proof

    Using the PDF of \(N_1\) above note that \[ G_1(n) = \sum_{k=1}^n [\log_b(k + 1) - \log_b(k)] = \log_b(n + 1), \quad n \in \{1, 2, \ldots, b - 1\} \] More generally, \( G_1(x) = G_1 \left(\lfloor x \rfloor\right) \) for \( x \in [1, b - 1] \)

    \( N_1 \) has quantile function \( G_1^{-1} \) given by \( G_1^{-1}(p) = \left\lceil b^p - 1\right\rceil \) for \( p \in (0, 1] \).

    1. The first quartile is \( \left\lceil b^{1/4} - 1 \right\rceil \).
    2. The median is \( \left\lceil b^{1/2} - 1 \right\rceil \).
    3. The third quartile is \( \left\lceil b^{3/4} - 1 \right\rceil \).
    Proof

    As usual, the formula for \( G_1^{-1}(p) \) follows from the CDF \(G\), by solving \( p = G(x) \) for \( x \) in terms of \( p \).

    Numerical values of the quantiles for the standard (base 10) distribution are given in an exercise below.

    Open the special distribution calculator and choose the Benford first digit distribution. Vary the base and note the shape and location of the distribution and probability density functions. For selected values of the base, compute the median and the first and third quartiles.

    For the most part the moments of \( N_1 \) do not have simple expressions. However, we do have the following result for the mean.

    \( \E(N_1) = (b - 1) - \log_b[(b - 1)!] \).

    Proof

    From the PDF of \(N_1\) above and using standard properties of the logarithm, \[ \E(N_1) = \sum_{n=1}^{b-1} n \log_b\left(\frac{n + 1}{n}\right) = \log_b\left[\prod_{n=1}^{b-1} \left(\frac{n + 1}{n}\right)^n\right] \] The product in the displayed equation simplifies to \( b^{b - 1} / (b - 1)! \), and the base \( b \) logarithm of this expression is \( (b - 1) - \log_b[(b - 1)!] \).

    Numerical values of the mean and variance for the standard (base 10) distribution are given in an exercise below.

    Opne the Special Distribution Simulator and select the Benford first digit distribution. Vary the base \( b \) with the input control and note the size and location of the mean \( \pm \) standard deviation bar. For various values of \( b \), run the simulation 1000 times and compare the empirical mean and standard deviation to the distribution mean and standard deviation..

    Since the quantile function has a simple, closed form, the Benford first digit distribution can be simulated via the random quantile method.

    Open the random quantile experiment and select the Benford first digit distribution. Vary the base \( b \) and note again the shape and location of the probability density function. For selected values of the base, run the experiment 1000 times and compare the empirical density function, mean, and standard deviation to their distributional counterparts.

    Higher Digits

    Now, to compute the joint probability density function of the first \(k\) significant digits, some additional notation will help.

    If \(n_1 \in \{1, 2, \ldots, b - 1\}\) and \(n_j \in \{0, 1, \ldots, b - 1\}\) for \(j \in \{2, 3, \ldots, k\}\), let \[ [n_1 \, n_2 \, \cdots \, n_k]_b = \sum_{j=1}^k n_j b ^{k - j} \]

    Of course, this is just the base \(b\) version of what we do in our standard base 10 system: we represent integers as strings of digits between 0 and 9 (except that the first digit cannot be 0). Here is a base 5 example: \[ [324]_5 = 3 \cdot 5^2 + 2 \cdot 5^1 + 4 \cdot 5^0 = 89 \]

    The joint probability density function \( h_k \) of \( (N_1, N_2, \ldots, N_k) \) is given by

    \[ h_k\left(n_1, n_2, \ldots, n_k\right) = \log_b \left( 1 + \frac{1}{[n_1 \, n_2 \, \cdots \, n_k]_b} \right), \quad n_1 \in \{1, 2, \ldots, b - 1\}, (n_2, \ldots, n_k) \in \{2, \ldots, b - 1\}^{k-1} \]
    Proof

    Note that \(\{N_1 = n_1, N_2 = n_2, \ldots, N_k = n_k\} = \{l \le Y \lt u \}\). where \[ l = \frac{[n_1 \, n_2 \, \cdots, n_k]_b}{b^k}, \; u = \frac{[n_1 \, n_2 \, \cdots \, n_k]_b + 1}{b^k} \] Hence using the PDF of \(Y\) and properties of logarithms, \[ h_k\left(n_1, n_2, \ldots, n_k\right) = \int_l^u \frac{1}{y \ln(b)} dy = \log_b(u) - \log_b(l) = \log_b\left([n_1 \, n_2 \, \cdots \, n_k]_b + 1\right) - \log_b\left([n_1 \, n_2 \, \cdots, n_k]_b\right) \]

    The probability density function of \( (N_1, N_2) \) in the standard (base 10) case is given in an exercise below. Of course, the probability density function of a given digit can be obtained by summing the joint probability density over the unwanted digits in the usual way. However, except for the first digit, these functions do not reduce to simple expressions.

    The probability density function \( g_2 \) of \( N_2 \) is given by \[ g_2(n) = \sum_{k=1}^{b-1} \log_b \left(1 + \frac{1}{[k \, n]_b} \right) = \sum_{k=1}^{b-1} \log_b \left(1 + \frac{1}{k \, b + n} \right), \quad n \in \{0, 1, \ldots, b - 1\} \]

    The probability density function of \( N_2 \) in the standard (base 10) case is given in an exercise below.

    Theoretical Explanation

    Aside from the empirical evidence noted by Newcomb and Benford (and many others since), why does Benford's law work? For a theoretical explanation, see the article A Statistical Derivation of the Significant Digit Law by Ted Hill.

    Computational Exercises

    In the following exercises, suppose that \( Y \) has the standard Benford mantissa distribution (the base 10 decimal case), and that \( \left(N_1, N_2, \ldots\right) \) are the digits of \( Y \).

    Find each of the following for the mantissa \( Y \)

    1. The density function \(f\).
    2. The mean and variance
    3. The quartiles
    Answer
    1. \(f(y) = \frac{1}{0.2303 y}, \quad y \in \left[\frac{1}{10}, 1\right)\)
    2. \(\E(Y) = 0.3909\), \(\var(Y) = 0.0622\)
    3. \( q_1 = 0.1778 \), \( q_2 = 0.3162 \), \( q_3 = 0.5623 \)

    For \( N_1 \), find each of the following numerically

    1. The probability density function
    2. The mean and variance
    3. The quartiles
    Answer
    1. \(n\) \(\P(N_1 = n)\)
      1 0.3010
      2 0.1761
      3 0.1249
      4 0.0969
      5 0.0792
      6 0.0669
      7 0.0580
      8 0.0512
      9 0.0458
    2. \(\E(N_1) = 3.4402\), \(\var(N_1) = 6.0567\)
    3. \( q_1 = 1 \), \( q_2 = 3 \), \( q_3 = 5 \)

    Explicitly compute the values of the joint probability density function of \((N_1, N_2)\).

    Answer
    \(\P(N_1 = n_1, N_2 = n_2)\) \(n_1 = 1\) 2 3 4 5 6 7 8 9
    \(n_2 = 0\) 0.0414 0.0212 0.0142 0.0107 0.0086 0.0072 0.0062 0.0054 0.0048
    1 0.0378 0.0202 0.0138 0.0105 0.0084 0.0071 0.0061 0.0053 0.0047
    2 0.0348 0.0193 0.0134 0.0102 0.0083 0.0069 0.0060 0.0053 0.0047
    3 0.0322 0.0185 0.0130 0.0100 0.0081 0.0068 0.0059 0.0052 0.0046
    4 0.0300 0.0177 0.0126 0.0098 0.0080 0.0067 0.0058 0.0051 0.0046
    5 0.0280 0.0170 0.0122 0.0092 0.0078 0.0066 0.0058 0.0051 0.0045
    6 0.0263 0.0164 0.0119 0.0093 0.0077 0.0065 0.0057 0.0050 0.0045
    7 0.0248 0.0158 0.0116 0.0091 0.0076 0.0064 0.0056 0.0050 0.0045
    8 0.0235 0.0152 0.0113 0.0090 0.0074 0.0063 0.0055 0.0049 0.0044
    9 0.0223 0.0147 0.0110 0.0088 0.0073 0.0062 0.0055 0.0049 0.0044

    For \( N_2 \), find each of the following numerically

    1. The probability density function
    2. \(\E(N_2)\)
    3. \(\var(N_2)\)
    Answer
    1. \(n\) \(\P(N_2 = n)\)
      0 0.1197
      1 0.1139
      2 0.1088
      3 0.1043
      4 0.1003
      5 0.0967
      6 0.0934
      7 0.0904
      8 0.0876
      9 0.0850
    2. \(\E(N_2) = 4.1847\)
    3. \(\var(N_2) = 0.8254\)

    Comparing the result for \(N_1\) and the result result for \(N_2\) , note that the distribution of \(N_2\) is flatter than the distribution of \(N_1\). In general, it turns out that distribution of \(N_k\) converges to the uniform distribution on \(\{0, 1, \ldots, b - 1\}\) as \(k \to \infty\). Interestingly, the digits are dependent.

    \(N_1\) and \(N_2\) are dependent.

    Proof

    This result follows from the joint PDF, the marginal PDF of \( N_1 \), and the marginal PDF of \( N_2 \) above.

    Find each of the following.

    1. \(\P(N_1 = 5, N_2 = 3, N_3 = 1)\)
    2. \(\P(N_1 = 3, N_2 = 1, N_3 = 5)\)
    3. \(\P(N_1 = 1, N_2 = 3, N_3 = 5)\)

    This page titled 5.39: Benford's Law is shared under a CC BY 2.0 license and was authored, remixed, and/or curated by Kyle Siegrist (Random Services) via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

    • Was this article helpful?