In mathematics, a geometric series is a series in which the ratio of successive adjacent terms is constant. In other words, the sum of consecutive terms of a geometric sequence forms a geometric series. Each term is therefore the geometric mean of its two neighbouring terms, similar to how the terms in an arithmetic series are the arithmetic means of their two neighbouring terms.

The geometric series 1/4 + 1/16 + 1/64 + 1/256 + ... shown as areas of purple squares. Each of the purple squares has 1/4 of the area of the next larger square (1/2×1/2 = 1/4, 1/4×1/4 = 1/16, etc.). The sum of the areas of the purple squares is one third of the area of the large square.
Another geometric series (coefficient a = 4/9 and common ratio r = 1/9) shown as areas of purple squares. The total purple area is S = a / (1 - r) = (4/9) / (1 - (1/9)) = 1/2, which can be confirmed by observing that the unit square is partitioned into an infinite number of L-shaped areas each with four purple squares and four yellow squares, which is half purple.

In general, a geometric series is written as , where is the initial term and is the common ratio between adjacent terms. For example, the series

is geometric because each successive term can be obtained by multiplying the previous term by .

Truncated geometric series are called "finite geometric series" in certain branches of mathematics, especially in 19th century calculus and in probability and statistics and their applications.

The standard generator form[1] expression for the infinite geometric series is

and the generator form expression for the finite geometric series is

Any finite geometric series has the sum , and when the infinite series converges to the value .

Geometric series have been studied in mathematics from at least the time of Euclid in his work, Elements, which explored geometric proportions. Archimedes further advanced the study through his work on infinite sums, particularly in calculating areas and volumes of geometric shapes (for instance calculating the area inside a parabola) and the early development of calculus. They serve as prototypes for frequently used mathematical tools such as Taylor series, Fourier series, and matrix exponentials.

Geometric series have been applied to model a wide variety of natural phenomena and social phenomena, such as the expansion of the universe where the common ratio is defined by Hubble's constant, the decay of radioactive carbon-14 atoms where the common ratio is defined by the half-life of carbon-14, probabilities of winning in games of chance where could be determined by the odds of a roulette wheel, and the economic values of investments where could be determined by a combination of inflation rates and interest rates.

Though geometric series are most commonly found and applied with the real or complex numbers for and , there are also important results and applications for matrix-valued geometric series, function-valued geometric series, p-adic number geometric series, and, most generally, geometric series of elements of abstract algebraic fields, rings, and semirings.

Parameters

edit

The geometric series   is an infinite series derived from a special type of sequence called a geometric progression, which is defined by just two parameters: the initial term   and the common ratio  . Finite geometric series   have a third parameter, the final term's power  

In applications with units of measurement, the initial term   provides the units of the series and the common ratio   is a dimensionless quantity.

The following table shows several geometric series with various initial terms and common ratios.

a r Example series
     
     ·
     
     
     
     
     
     

Initial term a

edit

The geometric series   has the same coefficient   in every term.[1] The first term of a geometric series is equal to this coefficient and is the parameter   of that geometric series, giving   its common interpretation: the "initial term."

This initial term defines the units of measurement of the series as a whole, if it has any, and in applications it will often be named according to a noun with those units. For instance   could be an "initial mass" in a radioactive decay problem, with units of mass of an isotope, an "initial payment" in mathematical finance, with units of some type of currency, or an "initial population" in demography or ecology, with units of a type such as nationality or species.

In generator form,   this term is technically written   instead of the bare  . This is equivalent because   for any number  

In contrast, a general power series   would have coefficients   that could vary from term to term. In other words, the geometric series is a special case of the power series. Connections between power series and geometric series are discussed below in the section § Connections to power series.

Common ratio r

edit

The parameter   is called the common ratio because it is the ratio of any term with the previous term in the series.

 

where   represents the  -th-power term of the geometric series.

The common ratio   can be thought of as a multiplier used to calculate each next term in the series from the previous term. It must be a dimensionless quantity.

When   it is often called a growth rate or rate of expansion and when   it is often called a decay rate or shrink rate, where the idea that it is a "rate" comes from interpreting   as a sort of discrete time variable. When an application area has specialized vocabulary for specific types of growth, expansion, shrinkage, and decay, that vocabulary will also often be used to name   parameters of geometric series. In economics, for instance, rates of increase and decrease of price levels are called inflation rates and deflation rates, while rates of increase in values of investments include rates of return and interest rates.

The interpretation of   as a time variable is often exactly correct in applications, such as the examples of amortized analysis of algorithmic complexity and calculating the present value of an annuity in § Applications below. In such applications it is also common to report a "growth rate"   in terms of another expression such as  , which is a percentage growth rate, or  , which is a doubling time, the opposite of a half-life.

Complex common ratio

edit
Complex geometric series (coefficient a = 1 and common ratio r = 0.5 e0t) converging to trace a circle. In the animation, each term of the geometric series is drawn as a vector twice: once at the origin and again within the head-to-tail vector summation that converges to the circle. The circle intersects the real axis at 2 (= 1/(1-1/2) when ω0t = 0) and at 2/3 (= 1/(1-(-1/2)) when ω0t = π).

The common ratio   can also be a complex number given by  , where   is the magnitude of the number as a vector in the complex plane,   is the angle or orientation of that vector,   is Euler's number, and  . In this case, the expanded form of the geometric series is

 

An example of how this behaves for   values that increase linearly over time with a constant angular frequency  , such that   is shown in the adjacent video. For   the geometric series becomes

 

where the first term is a vector of length   that does not change orientation and all the following terms are vectors of proportional lengths rotating in the complex plane at integer multiples of the fundamental angular frequency  , also known as harmonics of  . As the video shows, these sums trace a circle. The period of rotation around the circle is  .

Convergence

edit
 
The convergence of the geometric series with r=1/2 and a=1/2
 
The convergence of the geometric series with r=1/2 and a=1
 
Close-up view of the geometric series' partial sums over the range -1 < r < -0.5 as the first 11 terms of the geometric series 1 + r + r2 + r3 + ... are added, demonstrating alternating convergence. The partial sums' convergence limit 1 / (1 - r) is shown by the red dashed line.

The convergence of the sequence of partial sums of the geometric series depends on the magnitude of the common ratio   alone:

  • If  , the terms of the series approach zero (becoming smaller and smaller in magnitude) and the sequence of partial sums converge to a limit value of   proof is provided in the § Sum section.
  • If  , the terms of the series become larger and larger in magnitude and the sum of the terms also gets larger and larger in magnitude, so the series diverges.
  • If  , the sequence of partial sums of the series does not converge. When  , all the terms of the series are the same and the series grows to infinity. When  , the terms take two values alternately and therefore, the sequence of partial sums of the terms oscillates between two values. Consider, for example, Grandi's series:  . Partial sums of the terms oscillate between 1 and 0. Thus, the sequence of partial sums does not converge to any finite value. When   and  , the partial sums circulate periodically among the values  , never converging to a limit, and generally when   for any integer   and with any  , the partial sums of the series will circulate indefinitely with a period of  , never converging to a limit.

When the series converges, the rate of convergence and the pattern of convergence depend on the value of the common ratio  . The rate of convergence gets slower as   approaches  , see § Rate of convergence. If   and  , adjacent terms in the geometric series alternate between positive and negative and the partial sums of the terms oscillate above and below their eventual limit, whereas if   and   then terms all share the same sign and the partial sums of the terms approach their eventual limit monotonically.

For convenience, in this section the sum of the geometric series will be denoted by   and its partial sums (the sums of the series going up to only the nth power term) will be denoted  

The partial sum of the first   terms of a geometric series, up to and including the   term,

 

is given by the closed form

 

where r is the common ratio. The case   is just simple addition, a case of an arithmetic series. The formula for the partial sums   with   can be found as follows:[2][3][4]  

As   approaches 1, polynomial division or L'Hospital's rule recovers the case  .

 
Proof without words of the formula for the sum of a geometric series – if |r| < 1 and n → ∞, the r n term vanishes, leaving S = a/1 − r. This figure uses a slightly different convention for Sn than the main text, shifted by one term.

As   approaches infinity, the absolute value of r must be less than one for the series to converge, and when it does, the series converges absolutely. The sum of the infinite series then becomes

 

The formula holds for real   and for complex  .

This convergence result is widely applied to prove the convergence of other series as well, whenever those series's terms can be bounded from above by a suitable geometric series; that proof strategy is the basis for the general ratio test for the convergence of infinite series.

The question of whether an infinite sequence converges to a hypothetical limit depends essentially on the metric for distance between successive sequence values and the possible limit. In the above derivation, the metric for the distance between two values was implicitly the Euclidean distance between their locations on the number line or on the complex plane. However, the p-adic absolute value, which has become a critical notion in modern number theory, offers a metric for distance such that the geometric series 1 + 2 + 4 + 8 + ... with   and   actually does converge to   in the p-adic numbers using the p-adic absolute value metric even though   is outside the typical convergence range   on the number line according to a Euclidean metric.

Rate of convergence

edit

For any sequence  , its rate of convergence to a limit value   is determined by the parameters   and   such that

 

  is called the order of convergence, while   is called the rate of convergence. In the case of the sequence of partial sums of the geometric series, the relevant sequence is   and its limit is  . Therefore, the rate and order are found via

 

Using   and setting   gives

 

so the order of convergence of the geometric series is 1 and its rate of convergence is  . Convergence of order one is also called linear convergence or exponential convergence, depending on context.

Geometric proofs of convergence

edit
 
As an alternative to the algebraic derivation of the geometric series closed form formula, there is also the following geometric derivation. (TOP) Represent the first n+1 terms of S/a as areas of overlapping similar triangles.[5] For example, the area of the biggest overlapping (red) triangle is bh/2 = 2*1/2 = 1, which is the value of the first term of the geometric series. The area of the second biggest overlapping (green) triangle is bh/2 = (2r1/2)(r1/2)/2 = r, which is the value of the second term of the geometric series. Each progressively smaller triangle has its base and height scaled down by another factor of r1/2, resulting in a sequence of triangle areas 1, r, r2, r3, ... which is equal to the sequence of terms in the normalized geometric series S/a. (MIDDLE) In the order from largest to smallest, remove each triangle's overlapped area, which is always a fraction r of its area, and scale the remaining 1−r of the triangle's non-overlapped area by 1/(1−r) so the area of the formerly overlapped triangle, now the area of a non-overlapping trapezoid, remains the same. (BOTTOM) Aggregate the resulting n+1 non-overlapping trapezoids into a single non-overlapping trapezoid and calculate its area. The area of that aggregated trapezoid represents the value of the partial series. That area is equal to the outermost triangle minus the empty triangle tip: Sn/a = (1−rn+1) / (1−r), which approaches the limit S/a = 1/(1−r) when n approaches infinity and |r| < 1.

Alternatively, a geometric interpretation of the convergence for   is shown in the adjacent diagram. The area of the white triangle is the series remainder

 

Each additional term in the partial series reduces the area of that white triangle remainder by the area of the trapezoid representing the added term. The trapezoid areas (i.e., the values of the terms) get progressively thinner and shorter and closer to the origin. As the number of trapezoids approaches infinity, the white triangle remainder will vanish and therefore   will converge to  .

 
A geometric interpretation for the same case of common ratio r>1. (TOP) Represent the terms of a geometric series as the areas of overlapping similar triangles. (MIDDLE) From the largest to the smallest triangle, remove the overlapped left area portion (1/r) from the non-overlapped right area portion (1-1/r = (r-1)/r) and scale that non-overlapping trapezoid by r/(r-1) so its area is the same as the area of the original overlapping triangle. (BOTTOM) Calculate the area of the aggregate trapezoid as the area of the large triangle less the area of the empty small triangle at the large triangle's left tip. The large triangle is the largest overlapped triangle scaled by r/(r-1). The empty small triangle started as a but that area was transformed into a non-overlapping scaled trapezoid leaving an empty portion (1/r). However, that empty triangle of area a/r must also be scaled by r/(r-1) so its shape matches the shape of all the non-overlapped scaled trapezoids. Therefore, Sn = area of large triangle - area of empty small triangle = arn+1/(r-1) - a/(r-1) = a(rn+1-1)/(r-1), which does not converge to a limit.

In contrast, with  , shown in the next adjacent figure, the trapezoid areas representing the terms of the series would instead get progressively wider and taller and farther from the origin, not converging to the origin as terms and also not converging in sum as a series.

 
Converging alternating geometric series with common ratio r = -1/2 and coefficient a = 1. (TOP) Alternating positive and negative areas. (MIDDLE) Gaps caused by addition of adjacent areas. (BOTTOM) Gaps filled by broadening and decreasing the heights of the separated trapezoids.

The next adjacent diagram provides a geometric interpretation of a converging alternating geometric series with   where the areas corresponding to the negative terms are shown below the x-axis. When each positive area is paired with its adjacent smaller negative area, the result is a series of non-overlapping trapezoids, separated by gaps.

To eliminate these gaps, broaden each trapezoid so that it spans the rightmost   of the original triangle area instead of just the rightmost  . At the same time, to ensure the areas of the trapezoids remain consistent during this transformation, a rescaling is necessary. The required scaling factor   can be derived from the equation:

 

Simplifying this gives:

 

where   Because  , this scaling factor decreases the heights of the trapezoids to fill the gaps.

After the gaps are removed, pairs of terms in the converging alternating geometric series form a new converging geometric series with a common ratio  , reflecting the pairing of terms. The rescaled coefficient   compensates for the gap-filling.

History

edit

Zeno of Elea (c.495 – c.430 BC)

edit

2,500 years ago, Greek mathematicians believed[6] that an infinitely long list of positive numbers must sum to infinity. Therefore, Zeno of Elea created a paradox when he demonstrated that in order to walk from one place to another, one must first walk half the distance there, and then half of the remaining distance, and half of that remaining distance, and so on, covering infinitely many intervals before arriving. In doing so, he partitioned a fixed distance into an infinitely long list of halved remaining distances, each of which has length greater than zero. Zeno's paradox revealed to the Greeks that their assumption about an infinitely long list of positive numbers needing to add up to infinity was incorrect.

Euclid of Alexandria (c.300 BC)

edit
 
Elements of Geometry, Book IX, Proposition 35. "If there is any multitude whatsoever of continually proportional numbers, and equal to the first is subtracted from the second and the last, then as the excess of the second to the first, so the excess of the last will be to all those before it."

Euclid's Elements of Geometry[7] Book IX, Proposition 35, proof (of the proposition in adjacent diagram's caption):

Let AA', BC, DD', EF be any multitude whatsoever of continuously proportional numbers, beginning from the least AA'. And let BG and FH, each equal to AA', have been subtracted from BC and EF. I say that as GC is to AA', so EH is to AA', BC, DD'.

For let FK be made equal to BC, and FL to DD'. And since FK is equal to BC, of which FH is equal to BG, the remainder HK is thus equal to the remainder GC. And since as EF is to DD', so DD' to BC, and BC to AA' [Prop. 7.13], and DD' equal to FL, and BC to FK, and AA' to FH, thus as EF is to FL, so LF to FK, and FK to FH. By separation, as EL to LF, so LK to FK, and KH to FH [Props. 7.11, 7.13]. And thus as one of the leading is to one of the following, so (the sum of) all of the leading to (the sum of) all of the following [Prop. 7.12]. Thus, as KH is to FH, so EL, LK, KH to LF, FK, HF. And KH equal to CG, and FH to AA', and LF, FK, HF to DD', BC, AA'. Thus, as CG is to AA', so EH to DD', BC, AA'. Thus, as the excess of the second is to the first, so is the excess of the last is to all those before it. The very thing it was required to show.

The terseness of Euclid's propositions and proofs may have been a necessity. As is, the Elements of Geometry is over 500 pages of propositions and proofs. Making copies of this popular textbook was labor intensive given that the printing press was not invented until 1440. And the book's popularity lasted a long time: as stated in the cited introduction to an English translation, Elements of Geometry "has the distinction of being the world's oldest continuously used mathematical textbook." So being very terse was being very practical. The proof of Proposition 35 in Book IX could have been even more compact if Euclid could have somehow avoided explicitly equating lengths of specific line segments from different terms in the series. For example, the contemporary notation for geometric series (i.e., a + ar + ar2 + ar3 + ... + arn) does not label specific portions of terms that are equal to each other.

Also in the cited introduction the editor comments,

Most of the theorems appearing in the Elements were not discovered by Euclid himself, but were the work of earlier Greek mathematicians such as Pythagoras (and his school), Hippocrates of Chios, Theaetetus of Athens, and Eudoxus of Cnidos. However, Euclid is generally credited with arranging these theorems in a logical manner, so as to demonstrate (admittedly, not always with the rigour demanded by modern mathematics) that they necessarily follow from five simple axioms. Euclid is also credited with devising a number of particularly ingenious proofs of previously discovered theorems (e.g., Theorem 48 in Book 1).

To help translate the proposition and proof into a form that uses current notation, a couple modifications are in the diagram. First, the four horizontal line lengths representing the values of the first four terms of a geometric series are now labeled a, ar, ar2, ar3 in the diagram's left margin. Second, new labels A' and D' are now on the first and third lines so that all the diagram's line segment names consistently specify the segment's starting point and ending point.

Here is a phrase by phrase interpretation of the proposition:

Proposition in contemporary notation
"If there is any multitude whatsoever of continually proportional numbers" Taking the first n+1 terms of a geometric series Sn = a + ar + ar2 + ar3 + ... + arn
"and equal to the first is subtracted from the second and the last" and subtracting a from ar and arn
"then as the excess of the second to the first, so the excess of the last will be to all those before it." then (ar-a) / a = (arn-a) / (a + ar + ar2 + ar3 + ... + arn-1) = (arn-a) / Sn-1, which can be rearranged to the more familiar form Sn-1 = a(rn-1) / (r-1).

Similarly, here is a sentence by sentence interpretation of the proof:

Proof in contemporary notation
"Let AA', BC, DD', EF be any multitude whatsoever of continuously proportional numbers, beginning from the least AA'." Consider the first n+1 terms of a geometric series Sn = a + ar + ar2 + ar3 + ... + arn for the case r>1 and n=3.
"And let BG and FH, each equal to AA', have been subtracted from BC and EF." Subtract a from ar and ar3.
"I say that as GC is to AA', so EH is to AA', BC, DD'." I say that (ar-a) / a = (ar3-a) / (a + ar + ar2).
"For let FK be made equal to BC, and FL to DD'."
"And since FK is equal to BC, of which FH is equal to BG, the remainder HK is thus equal to the remainder GC."
"And since as EF is to DD', so DD' to BC, and BC to AA' [Prop. 7.13], and DD' equal to FL, and BC to FK, and AA' to FH, thus as EF is to FL, so LF to FK, and FK to FH."
"By separation, as EL to LF, so LK to FK, and KH to FH [Props. 7.11, 7.13]." By separation, (ar3-ar2) / ar2 = (ar2-ar) / ar = (ar-a) / a = r-1.
"And thus as one of the leading is to one of the following, so (the sum of) all of the leading to (the sum of) all of the following [Prop. 7.12]." The sum of those numerators and the sum of those denominators form the same proportion: ((ar3-ar2) + (ar2-ar) + (ar-a)) / (ar2 + ar + a) = r-1.
"And thus as one of the leading is to one of the following, so (the sum of) all of the leading to (the sum of) all of the following [Prop. 7.12]." And this sum of equal proportions can be extended beyond (ar3-ar2) / ar2 to include all the proportions up to (arn-arn-1) / arn-1.
"Thus, as KH is to FH, so EL, LK, KH to LF, FK, HF."
"And KH equal to CG, and FH to AA', and LF, FK, HF to DD', BC, AA'."
"Thus, as CG is to AA', so EH to DD', BC, AA'."
"Thus, as the excess of the second is to the first, so is the excess of the last is to all those before it." Thus, (ar-a) / a = (ar3-a) / S2. Or more generally, (ar-a) / a = (arn-a) / Sn-1, which can be rearranged in the more common form Sn-1 = a(rn-1) / (r-1).
"The very thing it was required to show." Q.E.D.

Archimedes of Syracuse (c.287 – c.212 BC)

edit
 
Archimedes' dissection of a parabolic segment into infinitely many triangles

Archimedes used the sum of a geometric series to compute the area enclosed by a parabola and a straight line. Archimedes' theorem states that the total area under the parabola is 4/3 of the area of the blue triangle. His method was to dissect the area into an infinite number of triangles as shown in the adjacent figure.

Archimedes determined that each green triangle has 1/8 the area of the blue triangle, each yellow triangle has 1/8 the area of a green triangle, and so forth.

Assuming that the blue triangle has area 1, the total area is an infinite series

 

The first term represents the area of the blue triangle, the second term the areas of the two green triangles, the third term the areas of the four yellow triangles, and so on. Simplifying the fractions gives

 

This is a geometric series with common ratio   and its sum is

 

This computation is an example of the method of exhaustion, an early version of integration. Using calculus, the same area could be found by a definite integral.

Nicole Oresme (c.1323 – 1382)

edit
 
A two dimensional geometric series diagram Nicole Oresme used to determine that the infinite series 1/2 + 2/4 + 3/8 + 4/16 + 5/32 + 6/64 + 7/128 + ... converges to 2.

In addition to his elegantly simple proof of the divergence of the harmonic series, Nicole Oresme[8] proved that the series

 

His diagram for his geometric proof, similar to the adjacent diagram, shows a two dimensional geometric series.

The first dimension is horizontal, in the bottom row, representing the geometric series with initial value   and common ratio  

 

The second dimension is vertical, where the bottom row is a new initial term   and each subsequent row above it shrinks according to the same common ratio  , making another geometric series with sum  ,

 

Although difficult to visualize beyond three dimensions, Oresme's insight generalizes to any dimension  . Denoting the sum of the  -dimensional series  , then using the limit of the  -dimensional geometric series,   as the initial term of a geometric series with the same common ratio in the next dimension, results in a recursive formula for   with the base case   given by the usual sum formula with an initial term  , so that:

 

within the range  , with   and   in Oresme's particular example.

Pascal's triangle exhibits the coefficients of these multi-dimensional geometric series,

 
        (closed form)   (expanded form)
     
     
     
     
     

where, as usual, the series converge to these closed forms only when  .

Examples

edit
  • Grandi's series – Infinite series summing alternating 1 and -1 terms: 1 − 1 + 1 − 1 + ⋯
  • 1 + 2 + 4 + 8 + ⋯ – Infinite series
  • 1 − 2 + 4 − 8 + ⋯ – infinite series
  • 1/2 + 1/4 + 1/8 + 1/16 + ⋯ – Mathematical infinite series
  • 1/2 − 1/4 + 1/8 − 1/16 + ⋯ – mathematical infinite series
  • 1/4 + 1/16 + 1/64 + 1/256 + ⋯ – Infinite series equal to 1/3 at its limit
  • A geometric series is a unit series (the series sum converges to one) if and only if |r| < 1 and a + r = 1 (equivalent to the more familiar form S = a / (1 - r) = 1 when |r| < 1). Therefore, an alternating series is also a unit series when -1 < r < 0 and a + r = 1 (for example, coefficient a = 1.7 and common ratio r = -0.7).
  • The terms of a geometric series are also the terms of a generalized Fibonacci sequence (Fn = Fn-1 + Fn-2 but without requiring F0 = 0 and F1 = 1) when a geometric series common ratio r satisfies the constraint 1 + r = r2, which according to the quadratic formula is when the common ratio r equals the golden ratio (i.e., common ratio r = (1 ± √5)/2).
  • The only geometric series that is a unit series and also has terms of a generalized Fibonacci sequence has the golden ratio as its coefficient a and the conjugate golden ratio as its common ratio r (i.e., a = (1 + √5)/2 and r = (1 - √5)/2). It is a unit series because a + r = 1 and |r| < 1, it is a generalized Fibonacci sequence because 1 + r = r2, and it is an alternating series because r < 0.
Repeating decimals and binaries
edit

Decimal numbers that have repeated patterns that continue forever, for instance     or   can be interpreted as geometric series and thereby converted to expressions of the ratio of two integers. For example, the repeated decimal fraction   can be written as the geometric series

 

where the initial term is   and the common ratio is  . The geometric series formula provides the integer ratio that corresponds to the repeating decimal:

 

An example that has four digits is the repeating decimal pattern,   This can be written as the geometric series

 

with initial term   and common ratio   The geometric series formula provides an integer ratio that corresponds to the repeating decimal:

 

This approach extends beyond repeating decimals, that is, base ten, to repeating patterns in other bases such as binary, that is, base two. For example, the binary representation of the number   is   where the binary pattern 110001 repeats indefinitely. That binary representation can be written as a geometric series of binary terms,

 

where the initial term is   expressed in base two   in base ten and the common ratio is   in base two   in base ten. Using the geometric series formula as before,

 

Applications

edit

Economics

edit

In economics, specifically in mathematical finance, geometric series are used to represent the present values of perpetual annuities (sums of money to be paid at regular intervals indefinitely into the future).

For example, suppose that a payment of $100 will be made to the owner of the perpetual annuity once per year (at the end of the year). In one simple model of the present value of future money, receiving $100 a year from now is worth less than an immediate $100 if one could invest the money now at a favorable interest rate. In particular, in that case, given a positive yearly interest rate  , the cost of an investment that produces $100 in the future is just   today, so the present value of $100 one year in the future is   today. More complex models of present value might account for the relative purchasing power of money today and in the future or account for changing personal utilities for having money now and in the future.

Continuing with the simple model and assuming a constant interest rate, a payment of $100 two years in the future would have a present value of   (squared because two years' worth of interest is lost by not receiving the money right now). Continuing that line of reasoning, the present value of receiving $100 per year in perpetuity would be

 

which is the infinite series:

 

This is a geometric series with common ratio   The sum is the first term divided by (one minus the common ratio):

 

For example, if the yearly interest rate is 10%   then the entire annuity has an estimated present value of  

This sort of calculation is used to compute the APR of a loan (such as a mortgage loan). It can also be used to estimate the present value of expected stock dividends, or the terminal value of a financial asset assuming a stable growth rate. However, the assumption that interest rates are constant is often incorrect and the payments are unlikely to in fact continue forever since the issuer of the perpetual annuity may lose its ability to make continued payments, so the estimates must be used with caution.

Computer science

edit

Fractal geometry

edit
The interior of the Koch snowflake is a union of infinitely many triangles. In the study of fractals, geometric series often arise as the perimeter, area, or volume of a self-similar figure.
If |r| < 1, representing each geometric series term as the area of a similar triangle creates a fractal in which more zooming in always reveals more similar triangles.

The area inside the Koch snowflake can be described as the union of infinitely many equilateral triangles (see figure). Each side of the green triangle is exactly 1/3 the size of a side of the large blue triangle, and therefore has exactly 1/9 the area. Similarly, each yellow triangle has 1/9 the area of a green triangle, and so forth. Taking the blue triangle as a unit of area, the total area of the snowflake is

 

The first term of this series represents the area of the blue triangle, the second term the total area of the three green triangles, the third term the total area of the twelve yellow triangles, and so forth. Excluding the initial 1, this series is geometric with constant ratio r = 4/9. The first term of the geometric series is a = 3(1/9) = 1/3, so the sum is

 

Thus the Koch snowflake has 8/5 of the area of the base triangle.

Trigonometric power series

edit

The Taylor series expansion of the arctangent function around zero, called the arctangent series, has been an important means for making approximate calculations in astronomy and optics for hundreds of years. It is traditionally called Gregory's series in Europe after the Scottish astronomer and mathematician James Gregory (1638 – 1675) though it is today more commonly attributed to the Keralan astronomer and mathematician Madhava of Sangamagrama (c. 1340 – c. 1425). It can be derived using differentiation, integration, and the sum of a geometric series.

The derivative of   is known to be  . This is a standard result derived as follows. Let   and   represent   and  ,[9]

 

Therefore, letting the arctan function equal the integral This is the power series expansion of the arctangent function.

Connections to power series

edit
 
Animation showing the convergence of the sequence of partial sums of the geometric series   as a functions of x (animated red lines) to their limit function in the domain  ,   (fixed blue line), a case of power series expansion about zero that converges only within the finite radius of convergence  .

Like the geometric series, a power series   has one parameter for a common variable raised to successive powers, denoted   here, corresponding to the geometric series's r, but it has additional parameters   one for each term in the series, for the distinct coefficients of each  , rather than just a single additional parameter   for all terms, the common coefficient of   in each term of a geometric series.

The geometric series can therefore be considered a class of power series in which the sequence of coefficients satisfies   for all   and  . This special class of power series plays an important role in mathematics, for instance for the study of ordinary generating functions in combinatorics and the summation of divergent series in analysis. Many other power series can be written as transformations and combinations of geometric series, making the geometric series formula a convenient tool for calculating formulas for those power series as well.

As a power series, the geometric series has a radius of convergence of 1. This could be seen as a consequence of the Cauchy–Hadamard theorem and the fact that   for any   or as a consequence of the ratio test for the convergence of infinite series, with   implying convergence only for   However, both the ratio test and the Cauchy–Hadamard theorem are proven using the geometric series formula as a logically prior result, so such reasoning would be subtly circular.

Derivations of other power series formulas

edit

Infinite series formulas

edit

One can use simple variable substitutions to calculate some useful closed form infinite series formulas. For an infinite series containing only even powers of  , for instance,   and for odd powers only,  

In cases where the sum does not start at k = 0, one can use a shift of the index of summation together with a variable substitution,   The formulas given above are strictly valid only for |r| < 1. The latter formula is valid in every Banach algebra, as long as the norm of r is less than one, and also in the field of p-adic numbers if |r|p < 1.

One can also differentiate to calculate formulas for related sums. For example,  

This formula is only strictly valid for |r| < 1 as well. From similar derivations, it follows that, for |r| < 1,  

It is also possible to use complex geometric series to calculate the sums of some trigonometric series using complex exponentials and Euler's formula. For example, consider the proposition  

This can be proven via the fact that  Substituting this into the original series gives  

This is the difference of two geometric series with common ratios   and  , and so the proof of the original proposition follows via two straightforward applications of the formula for infinite geometric series and then rearrangement of the result using  and   to complete the proof.

Finite series formulas

Like for the infinite series, one can use variable substitutions and changes of the index of summation to derive other finite power series formulas from the finite geometric series formulas. If one were to begin the sum not from k=1 or 0 but from a different value, say  , then 

For a geometric series containing only even powers of  , either multiply the finite sum by   to derive a closed form for the partial sums when  : 

or, equivalently, take   as the common ratio   and use the standard formula.

For a series with only odd powers of  , the same derivation via multiplying by   applies when  :  or, equivalently, take   for   and   for   in the standard form.

Differentiating such formulas with respect to   can give the formulas

 

For example:

 

An exact formula for any of the generalized sums   when   is

 

where   denotes a Stirling number of the second kind.[10]

Generalizations beyond real and complex values

edit

While geometric series with real and complex number parameters   and   are most common, geometric series of more general terms such as functions, matrices, and p-adic numbers also find application. The mathematical operations used to express a geometric series given its parameters are simply addition and repeated multiplication, and so it is natural, in the context of modern algebra, to define geometric series with parameters from any ring or field. Further generalization to geometric series with parameters from semirings is more unusual, but also has applications, for instance in the study of fixed-point iteration of transformation functions.

In order to analyze the convergence of these general geometric series, then on top of addition and multiplication, one must also have some metric of distance between partial sums of the series. This can introduce new subtleties into the questions of convergence, such as the distinctions between uniform convergence and pointwise convergence in series of functions, and can lead to strong contrasts with intuitions from the real numbers, such as in the convergence of the series 1 + 2 + 4 + 8 + ... with   and   to   in the p-adic numbers using the p-adic absolute value.

When the multiplication of the parameters is not commutative, as it often is not for matrices or general physical operators, particularly in quantum mechanics, then the standard way of writing the geometric series,  , multiplying from the right, may need to be distinguished from the alternative  , multiplying from the left, and also the symmetric  , multiplying half on each side. These choices may correspond to important alternatives with different strengths and weaknesses in applications, as in the case of ordering the mutual interferences of drift and diffusion differently at infinitesimal temporal scales in Ito integration and Stratonovitch integration in stochastic calculus.

See also

edit

Notes

edit
  1. ^ a b Riddle, Douglas F. Calculus and Analytic Geometry, Second Edition Belmont, California, Wadsworth Publishing, p. 566, 1970.
  2. ^ Abramowitz & Stegun (1972, p. 10)
  3. ^ Moise (1967, p. 48)
  4. ^ Protter & Morrey (1970, pp. 639–640)
  5. ^ Hairer E.; Wanner G. (1996). Analysis by Its History. Springer. p. 188. Section III.2, Figure 2.1
  6. ^ Riddle, Douglas E (1974). Calculus and Analytic Geometry (2nd ed.). Wadsworth Publishing. p. 556. ISBN 053400301-X.
  7. ^ Euclid; J.L. Heiberg (2007). Euclid's Elements of Geometry (PDF). Translated by Richard Fitzpatrick. Richard Fitzpatrick. ISBN 978-0615179841. Archived (PDF) from the original on 2013-08-11.
  8. ^ Babb, J (2003). "Mathematical Concepts and Proofs from Nicole Oresme: Using the History of Calculus to Teach Mathematics" (PDF). Winnipeg: The Seventh International History, Philosophy and Science Teaching conference. pp. 11–12, 21. Archived (PDF) from the original on 2021-05-27.
  9. ^ Riddle, Douglas (1974). Calculus and Analytic Geometry (second ed.). California: Wadsworth Publishing. p. 310. ISBN 0-534--00301-X.
  10. ^ "Set Partitions: Stirling Numbers". Digital Library of Mathematical Functions. Retrieved 24 May 2018.

References

edit
  • Abramowitz, M.; Stegun, I. A., eds. (1972). Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables (9th printing ed.). New York: Dover. p. 10.
  • Andrews, George E. (1998). "The geometric series in calculus". The American Mathematical Monthly. 105 (1). Mathematical Association of America: 36–40. doi:10.2307/2589524. JSTOR 2589524.
  • Arfken, G. Mathematical Methods for Physicists, 3rd ed. Orlando, FL: Academic Press, pp. 278–279, 1985.
  • Beyer, W. H. CRC Standard Mathematical Tables, 28th ed. Boca Raton, FL: CRC Press, p. 8, 1987.
  • Courant, R. and Robbins, H. "The Geometric Progression." §1.2.3 in What Is Mathematics?: An Elementary Approach to Ideas and Methods, 2nd ed. Oxford, England: Oxford University Press, pp. 13–14, 1996.
  • Hall, Brian C. (2015), Lie groups, Lie algebras, and representations: An elementary introduction, Graduate Texts in Mathematics, vol. 222 (2nd ed.), Springer, ISBN 978-3-319-13466-6
  • Horn, Roger A.; Johnson, Charles R. (1990). Matrix Analysis. Cambridge University Press. ISBN 978-0-521-38632-6..
  • James Stewart (2002). Calculus, 5th ed., Brooks Cole. ISBN 978-0-534-39339-7
  • Larson, Hostetler, and Edwards (2005). Calculus with Analytic Geometry, 8th ed., Houghton Mifflin Company. ISBN 978-0-618-50298-1
  • Moise, Edwin E. (1967), Calculus: Complete, Reading: Addison-Wesley
  • Pappas, T. "Perimeter, Area & the Infinite Series." The Joy of Mathematics. San Carlos, CA: Wide World Publ./Tetra, pp. 134–135, 1989.
  • Protter, Murray H.; Morrey, Charles B. Jr. (1970), College Calculus with Analytic Geometry (2nd ed.), Reading: Addison-Wesley, LCCN 76087042
  • Roger B. Nelsen (1997). Proofs without Words: Exercises in Visual Thinking, The Mathematical Association of America. ISBN 978-0-88385-700-7

History and philosophy

edit
  • C. H. Edwards Jr. (1994). The Historical Development of the Calculus, 3rd ed., Springer. ISBN 978-0-387-94313-8.
  • Swain, Gordon and Thomas Dence (April 1998). "Archimedes' Quadrature of the Parabola Revisited". Mathematics Magazine. 71 (2): 123–30. doi:10.2307/2691014. JSTOR 2691014.
  • Eli Maor (1991). To Infinity and Beyond: A Cultural History of the Infinite, Princeton University Press. ISBN 978-0-691-02511-7
  • Morr Lazerowitz (2000). The Structure of Metaphysics (International Library of Philosophy), Routledge. ISBN 978-0-415-22526-7

Economics

edit
  • Carl P. Simon and Lawrence Blume (1994). Mathematics for Economists, W. W. Norton & Company. ISBN 978-0-393-95733-4
  • Mike Rosser (2003). Basic Mathematics for Economists, 2nd ed., Routledge. ISBN 978-0-415-26784-7

Biology

edit
  • Edward Batschelet (1992). Introduction to Mathematics for Life Scientists, 3rd ed., Springer. ISBN 978-0-387-09648-3
  • Richard F. Burton (1998). Biology by Numbers: An Encouragement to Quantitative Thinking, Cambridge University Press. ISBN 978-0-521-57698-7

Computer science

edit
  • John Rast Hubbard (2000). Schaum's Outline of Theory and Problems of Data Structures With Java, McGraw-Hill. ISBN 978-0-07-137870-3
edit