Вы находитесь на странице: 1из 13

Geometric series

 Read in another language


 Download PDF
 Watch
 Edit
This article is about infinite geometric series. For finite sums, see geometric progression.

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.

In mathematics, a geometric series is a series with a constant ratio between


successive terms. For example, the series

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


previous term by 1/2.

Geometric series are among the simplest examples of infinite series with finite
sums, although not all of them have this property. Historically, geometric series
played an important role in the early development of calculus, and they continue
to be central in the study of convergence of series. Geometric series are used
throughout mathematics, and they have important applications
in physics, engineering, biology, economics, computer science, queueing theory,
and finance.
Contents

Common ratioEdit
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

The terms of a geometric series form a geometric progression, meaning that the
ratio of successive terms in the series is constant. This relationship allows for
the representation of a geometric series using only two terms, r and a. The
term r is the common ratio, and a is the first term of the series. As an
example the geometric series given in the introduction,

may simply be written as

, with and .

The following table shows several geometric series with different start terms and
common ratios:

Start term, a Common ratio, r Example series


4 10 4 + 40 + 400 + 4000 + 40,000 + ···
9 1/3 9 + 3 + 1 + 1/3 + 1/9 + ···
7 1/10 7 + 0.7 + 0.07 + 0.007 + 0.0007 + ···
3 1 3 + 3 + 3 + 3 + 3 + ···
1 −1/2 1 − 1/2 + 1/4 − 1/8 + 1/16 − 1/32 + ···
3 –1 3 − 3 + 3 − 3 + 3 − ···
The behavior of the terms depends on the common ratio r:
If r is between −1 and +1, the terms of the series approach zero in the
limit (becoming smaller and smaller in magnitude), and the series
converges to a sum. In the case above, where r is 1/2, the series
converges to 1.
If r is greater than one or less than minus one the terms of the series
become larger and larger in magnitude. The sum of the terms also gets
larger and larger, and the series has no sum. (The series diverges.)
If r is equal to one, all of the terms of the series are the same. The
series diverges.
If r is minus one the terms take two values alternately (e.g. 2, −2, 2,
−2, 2,... ). The sum of the terms oscillates between two values (e.g. 2,
0, 2, 0, 2,... ). This is a different type of divergence and again the
series has no sum. See for example Grandi's series: 1 − 1 + 1 − 1 + ···.

SumEdit
The sum of a geometric series is finite as long as the absolute value of the ratio
is less than 1; as the numbers near zero, they become insignificantly small,
allowing a sum to be calculated despite the series containing infinitely many
terms. The sum can be computed using the self-similarity of the series.
ExampleEdit

Visual derivation of the sum of infinite terms of a geometric series

Consider the sum of the following geometric series:

This series has common ratio 2/3. If we multiply through by this common ratio,
then the initial 1 becomes a 2/3, the 2/3 becomes a 4/9, and so on:
This new series is the same as the original, except that the first term is
missing. Subtracting the new series (2/3)s from the original series s cancels
every term in the original but the first,

A similar technique can be used to evaluate any self-similar expression.


FormulaEdit

For , the sum of the first n terms of a geometric series is

where a is the first term of the series, and r is the common ratio. We can
derive this formula as follows:

so,

As n goes to infinity, the absolute value of r must be less than one for the
series to converge. The sum then becomes

When a = 1, this can be simplified to

the left-hand side being a geometric series with common ratio r.


The formula also holds for complex r, with the corresponding restriction, the
modulus of r is strictly less than one.
Proof of convergenceEdit
We can prove that the geometric series converges using the sum formula for
a geometric progression:

Since (1 + r + r2 + ... + rn)(1−r) = 1−rn+1 and rn+1 → 0 for | r | < 1.


Convergence of geometric series can also be demonstrated by rewriting the series
as an equivalent telescoping series. Consider the function,
Note that

Thus,

If

then

So S converges to

ApplicationsEdit
Repeating decimalsEdit
Main article: Repeating decimal

A repeating decimal can be thought of as a geometric series whose common ratio


is a power of 1/10. For example:

The formula for the sum of a geometric series can be used to convert the
decimal to a fraction,

The formula works not only for a single repeating figure, but also for a repeating
group of figures. For example:

Note that every series of repeating consecutive decimals can be conveniently


simplified with the following:
That is, a repeating decimal with repeat length n is equal to the quotient of
the repeating part (as an integer) and 10n - 1.
Archimedes' quadrature of the parabolaEdit

Archimedes' dissection of a parabolic segment into infinitely many triangles

Main article: The Quadrature of the Parabola

Archimedes used the sum of a geometric series to compute the area enclosed by
a parabola and a straight line. His method was to dissect the area into an
infinite number of triangles.

Archimedes' Theorem states that the total area under the parabola is 4/3 of
the area of the blue triangle.

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 sum:

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 1/4 and the fractional part is equal
to

The sum is

This computation uses the method of exhaustion, an early version of integration.


Using calculus, the same area could be found by a definite integral.
Fractal geometryEdit

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.
For example, 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.

Zeno's paradoxesEdit
Main article: Zeno's paradoxes

The convergence of a geometric series reveals that a sum involving an infinite


number of summands can indeed be finite, and so allows one to resolve many
of Zeno's paradoxes. For example, Zeno's dichotomy paradox maintains that
movement is impossible, as one can divide any finite path into an infinite
number of steps wherein each step is taken to be half the remaining distance.
Zeno's mistake is in the assumption that the sum of an infinite number of
finite steps cannot be finite. This is of course not true, as evidenced by the

convergence of the geometric series with .


EuclidEdit
Book IX, Proposition 35[1] of Euclid's Elements expresses the partial sum of a
geometric series in terms of members of the series. It is equivalent to the
modern formula.
EconomicsEdit
Main article: Time value of money

In economics, geometric series are used to represent the present value of


an annuity (a sum of money to be paid in regular intervals).
For example, suppose that a payment of $100 will be made to the owner of
the annuity once per year (at the end of the year) in perpetuity. Receiving
$100 a year from now is worth less than an immediate $100, because one
cannot invest the money until one receives it. In particular, the present value of

$100 one year in the future is $100 / (1 + ), where is the yearly


interest rate.
Similarly, a payment of $100 two years in the future has a present value of

$100 / (1 + )2 (squared because two years' worth of interest is lost by not


receiving the money right now). Therefore, the present value of receiving $100
per year in perpetuity is

which is the infinite series:

This is a geometric series with common ratio 1 / (1 + ). The sum is the


first term divided by (one minus the common ratio):

For example, if the yearly interest rate is 10% ( = 0.10), then the entire
annuity has a present value of $100 / 0.10 = $1000.
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 security.
Geometric power seriesEdit
The formula for a geometric series

can be interpreted as a power series in the Taylor's theorem sense, converging

where . From this, one can extrapolate to obtain other power series. For
example,

By differentiating the geometric series, one obtains the variant[2]

Similarly obtained are:

and
See alsoEdit
 0.999...
 Asymptote
 Divergent geometric series
 Generalized hypergeometric function
 Geometric progression
 Neumann series
 Ratio test
 Root test
 Series (mathematics)

Specific geometric seriesEdit


 Grandi's series: 1 − 1 + 1 − 1 + ⋯
 1 + 2 + 4 + 8 + ⋯
 1 − 2 + 4 − 8 + ⋯
 1/2 + 1/4 + 1/8 + 1/16 + ⋯
 1/2 − 1/4 + 1/8 − 1/16 + ⋯
 1/4 + 1/16 + 1/64 + 1/256 + ⋯

ReferencesEdit
1. ^ "Euclid's Elements, Book IX, Proposition 35". Aleph0.clarku.edu.
Retrieved 2013-08-01.
2. ^ Taylor, Angus E. (1955). Advanced Calculus. Blaisdell. p. 603.

 Abramowitz, M. and Stegun, I. A. (Eds.). Handbook of Mathematical Functions with


Formulas, Graphs, and Mathematical Tables, 9th printing. New York: Dover, p. 10,
1972.

 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.

 Pappas, T. "Perimeter, Area & the Infinite Series." The Joy of Mathematics. San
Carlos, CA: Wide World Publ./Tetra, pp. 134–135, 1989.

 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
 Roger B. Nelsen (1997). Proofs without Words: Exercises in Visual Thinking, The
Mathematical Association of America. ISBN 978-0-88385-700-7
 Andrews, George E. (1998). "The geometric series in calculus". The American
Mathematical Monthly. Mathematical Association of America. 105 (1): 36–
40. doi:10.2307/2589524. JSTOR 2589524.

History and philosophyEdit


 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

EconomicsEdit
 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

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

External linksEdit
 Hazewinkel, Michiel, ed. (2001) [1994], "Geometric progression", Encyclopedia of
Mathematics, Springer Science+Business Media B.V. / Kluwer Academic
Publishers, ISBN 978-1-55608-010-4
 Weisstein, Eric W. "Geometric Series". MathWorld.
 Geometric Series at PlanetMath.org.
 Peppard, Kim. "College Algebra Tutorial on Geometric Sequences and Series".
West Texas A&M University.
 Casselman, Bill. "A Geometric Interpretation of the Geometric Series". Archived
from the original (Applet) on 2007-09-29.
 "Geometric Series" by Michael Schreiber, Wolfram Demonstrations Project,
2007.









Divergent geometric series
 Read in another language
 Download PDF
 Watch
 Edit
In mathematics, an infinite geometric series of the form

is divergent if and only if | r | ≥ 1. Methods for summation of divergent series


are sometimes useful, and usually evaluate divergent geometric series to a sum
that agrees with the formula for the convergent case

This is true of any summation method that possesses the properties


of regularity, linearity, and stability.
Contents

Вам также может понравиться