Fact-checked by Grok 2 weeks ago

Geometric series

A geometric series is a mathematical series formed by a sequence of terms where each term after the first is obtained by multiplying the preceding term by a fixed, nonzero constant known as the common ratio r. The general form of an infinite geometric series is \sum_{n=0}^{\infty} a r^n, where a is the first term. The sum of the first n terms of a finite geometric series, denoted S_n, is calculated using the formula S_n = a \frac{1 - r^n}{1 - r} for r \neq 1. For the infinite case, the series converges to the sum S = \frac{a}{1 - r} the of the common satisfies |r| < 1; otherwise, it diverges. This convergence criterion is a cornerstone of series analysis in calculus, distinguishing geometric series from other types like harmonic or p-series. Historically, geometric series trace their origins to ancient Greek mathematics, with early appearances in Zeno's paradoxes of the 5th century BCE, such as the Achilles and the tortoise dilemma, which implicitly relies on the summation of an infinite geometric series to resolve apparent contradictions in motion and infinity. These paradoxes highlighted the need for rigorous understanding of infinite sums, paving the way for later developments in analysis by mathematicians like Archimedes and, in the modern era, Isaac Newton and Gottfried Wilhelm Leibniz. Beyond pure mathematics, geometric series find extensive applications across disciplines due to their ability to model exponential growth and decay processes. In finance, they underpin calculations of compound interest, where periodic returns form a geometric progression. In physics and biology, they describe phenomena like radioactive decay and population growth under constant multiplication rates, and in engineering, signal attenuation. Additionally, in computer science, geometric series appear in analyzing the time complexity of recursive algorithms.

Definition and Fundamentals

Definition

A geometric series is the sum of the terms of a geometric progression, which is a sequence of numbers where each term after the first is found by multiplying the previous one by a fixed, non-zero constant called the common ratio. The underlying geometric progression takes the form a, ar, ar^2, ar^3, \dots, where a is the first term and r is the common ratio, with the general term given by ar^{k} for k = 0, 1, 2, \dots. For the finite case, a geometric series consists of a finite number of terms from this progression, expressed as the partial sum S_n = \sum_{k=0}^{n-1} ar^k, where n denotes the number of terms; equivalently, it may start the index at 1 as \sum_{k=1}^{n} ar^{k-1}. In the infinite case, the geometric series is the sum over infinitely many terms, S = \sum_{k=0}^{\infty} ar^k, or alternatively \sum_{k=1}^{\infty} ar^{k-1}, and is understood as the limit of the partial sums S_n as n approaches infinity, provided this limit exists. Notation for partial sums often uses S_n to emphasize the finite approximation to the infinite series.

Examples

A classic example of a finite geometric series is the sum of the first four terms of the sequence 1, 2, 4, 8, where the first term is 1 and the common ratio is 2; this series totals 15. To illustrate the terms and partial sums, consider the following table for this series:
Term numberTerm valuePartial sum
111
223
347
4815
An infinite geometric series is demonstrated by 1 + \frac{1}{2} + \frac{1}{4} + \frac{1}{8} + \cdots, with first term 1 and common ratio \frac{1}{2}, which converges to 2. The partial sums of this series approach the limit as follows:
Term numberTerm valuePartial sum
111
2\frac{1}{2}1.5
3\frac{1}{4}1.75
4\frac{1}{8}1.875
\dots\dotsapproaching 2
Repeating decimals provide an introductory real-world representation of infinite ; for instance, 0.333\dots equals \frac{1}{3} and can be written as the series \frac{3}{10} + \frac{3}{100} + \frac{3}{1000} + \cdots, with first term \frac{3}{10} and common ratio \frac{1}{10}. A common pitfall arises with series where the absolute value of the common ratio exceeds 1, such as -0.5 + 1.5 - 4.5 + \cdots with ratio -3, which diverges without a finite sum.

Finite Geometric Series

Sum formula

The sum S_n of the first n terms of a finite geometric series, with first term a and common ratio r \neq 1, is given by the closed-form expression S_n = a \frac{1 - r^n}{1 - r}. This formula allows for efficient computation of the total without adding each term individually. When r = 1, the series degenerates into an arithmetic sequence with constant terms, and the sum simplifies to S_n = n a. Key properties of the finite geometric sum include its independence from the order of term addition when all terms are positive, ensuring the total remains unchanged regardless of summation sequence. Additionally, the partial sums exhibit a telescoping nature, where multiplying the series by r leads to cancellations that isolate the first and last terms in the difference. For example, consider the series $3 + 6 + 12 + 24 + 48 (a = 3, r = 2, n = 5). Applying the formula yields S_5 = 3 \frac{1 - 2^5}{1 - 2} = 3 \frac{1 - 32}{-1} = 3 \times 31 = 93. This matches the direct addition: $3 + 6 + 12 + 24 + 48 = 93. Edge cases include r = 0, where the series is a, 0, 0, \dots, resulting in the trivial sum S_n = a for n \geq 1, as subsequent terms vanish. The case r = 1 was addressed earlier, highlighting the transition from geometric to arithmetic behavior.

Derivation of the sum

The sum S_n of the first n terms of a finite with first term a and common ratio r (where r \neq 1) is given by S_n = a + ar + ar^2 + \dots + ar^{n-1}. To derive the closed-form formula, multiply both sides by r: r S_n = ar + ar^2 + ar^3 + \dots + ar^n. Subtracting the original sum from this equation yields: S_n - r S_n = a - ar^n, which simplifies to S_n (1 - r) = a (1 - r^n). Solving for S_n gives S_n = \frac{a (1 - r^n)}{1 - r}. This algebraic manipulation, often attributed to early developments in series summation techniques, provides a direct way to compute the sum without explicit addition of terms. An alternative approach to establishing the formula uses mathematical induction on n. For the base case n = 1, S_1 = a, and the formula holds: \frac{a (1 - r^1)}{1 - r} = a. Assume the formula is true for n = k, so S_k = \frac{a (1 - r^k)}{1 - r}. For n = k+1, S_{k+1} = S_k + a r^k = \frac{a (1 - r^k)}{1 - r} + a r^k = \frac{a (1 - r^k) + a r^k (1 - r)}{1 - r} = \frac{a (1 - r^{k+1})}{1 - r}. By induction, the formula holds for all positive integers n. This proof confirms the algebraic result rigorously for discrete cases. Geometrically, the sum can be interpreted as the total area of a stepped figure where each step's height is scaled by r, forming a telescoping pattern upon multiplication and subtraction, akin to filling a region iteratively in a square. For instance, with a = 1/2 and r = 1/2, the partial sums represent successively larger shaded portions of a unit square approaching but not reaching full coverage. To verify, consider n=2: S_2 = a + ar = a(1 + r), and the formula gives \frac{a(1 - r^2)}{1 - r} = a(1 + r), matching exactly. For n=3: S_3 = a + ar + ar^2 = a(1 + r + r^2), and \frac{a(1 - r^3)}{1 - r} = a(1 + r + r^2), confirming correctness for small values.

Infinite Geometric Series

Convergence criteria

An infinite geometric series \sum_{k=0}^{\infty} ar^k, where a \neq 0 is the first term and r is the common ratio, converges if and only if |r| < 1, and diverges otherwise when |r| \geq 1. This condition ensures that the terms of the series decrease in magnitude over time, allowing the partial sums to approach a finite limit. The convergence criterion can be understood through the ratio test, a standard tool for analyzing series. For the general term u_k = ar^k, the ratio test examines \lim_{k \to \infty} \left| \frac{u_{k+1}}{u_k} \right| = |r|; the series converges absolutely if this limit is less than 1, which occurs precisely when |r| < 1. If the limit equals 1, the test is inconclusive, but for geometric series, this case corresponds to divergence as detailed below. When |r| < 1, the series converges to a finite value, with successive terms becoming arbitrarily small, so the partial sums S_n = \sum_{k=0}^{n} ar^k stabilize toward a limit as n increases indefinitely. In contrast, if |r| > 1, the terms grow without bound in , causing the partial sums to diverge to \infty or -\infty depending on the sign of a and r. For |r| = 1, the series diverges: if r = 1, it becomes a constant series \sum a that sums to infinity; if r = -1, it alternates between a and -a, yielding partial sums that oscillate without settling. The behavior of partial sums S_n illustrates convergence graphically: for |r| < 1, plotting S_n against n shows a curve approaching a horizontal asymptote, reflecting the series' finite limit, whereas for |r| \geq 1, the plot either rises unbounded or oscillates, confirming divergence. This visual progression underscores how finite geometric sums serve as foundational steps toward understanding infinite .

Sum formula and proof

The sum of an infinite geometric series with first term a and common ratio r, where |r| < 1, is given by the formula S = \sum_{k=0}^{\infty} a r^k = \frac{a}{1 - r}. One proof derives this by taking the limit of the partial sums from the finite geometric series. The partial sum up to n terms is S_n = a \frac{1 - r^{n+1}}{1 - r}. As n \to \infty, since |r| < 1, it follows that r^{n+1} \to 0, so S = \lim_{n \to \infty} S_n = \frac{a}{1 - r}. An alternative proof uses the series equation directly. Let S = a + a r + a r^2 + \cdots. Multiplying by r gives r S = a r + a r^2 + a r^3 + \cdots. Subtracting yields S - r S = a, so S (1 - r) = a, and thus S = \frac{a}{1 - r}, provided |r| < 1. For approximations, the error when using the partial sum S_n is the remainder R_n = S - S_n = a r^{n+1} / (1 - r), so the absolute error bound is |R_n| = |a| |r|^{n+1} / |1 - r|. This bound decreases exponentially as n increases when |r| < 1. This formula extends to complex numbers, where the series converges to a / (1 - r) inside the open unit disk |r| < 1 in the complex plane.

Mathematical Connections

Relation to power series

The geometric series serves as a foundational example of a power series, representing the function \frac{1}{1 - x} as the infinite sum \sum_{k=0}^{\infty} x^k for |x| < 1, where the first term a = 1 and the common ratio r = x. This expansion illustrates how a simple rational function can be expressed as a power series centered at x = 0, with all coefficients equal to 1, highlighting the series' role in approximating analytic functions within its interval of convergence. The radius of convergence for this series is 1, determined via the ratio test, which examines \lim_{n \to \infty} \left| \frac{a_{n+1}}{a_n} \right| = |x|; the series converges when this limit is less than 1, i.e., |x| < 1, and diverges for |x| \geq 1. The singularity at x = 1 marks the boundary of convergence, where the partial sums approach infinity, underscoring the series' analytic properties in the complex plane. In the broader context of power series, the geometric series is a special case of the general form \sum_{k=0}^{\infty} c_k (x - a)^k, where the coefficients c_k = 1 for all k and the center a = 0, demonstrating uniform convergence on compact subsets within the disk of convergence. Historically, Isaac Newton employed the geometric series to generalize the binomial theorem, extending it to non-integer exponents by recognizing patterns in infinite expansions, such as for (1 + x)^n where n is fractional, which laid groundwork for modern power series techniques. In calculus applications, this approach parallels the derivation of Taylor series for functions like \ln(1 + x), obtained by integrating the geometric series for \frac{1}{1 + t} = \sum_{k=0}^{\infty} (-1)^k t^k from 0 to x, yielding \ln(1 + x) = \sum_{k=1}^{\infty} (-1)^{k+1} \frac{x^k}{k} for |x| < 1; similarly, the Taylor series for e^x = \sum_{k=0}^{\infty} \frac{x^k}{k!} is developed through repeated differentiation, illustrating the power series method's versatility.

Role in generating functions

In enumerative combinatorics, the geometric series forms the foundational ordinary generating function (OGF) for sequences exhibiting geometric growth. Specifically, the OGF for the sequence a_k = r^k (for k \geq 0) is \sum_{k=0}^\infty r^k x^k = \frac{1}{1 - r x}, treated as a formal power series without regard to convergence or analytically within |r x| < 1. This structure arises naturally in counting unlabeled combinatorial objects, such as the total number of binary strings of length n, which is $2^n; the corresponding OGF is \frac{1}{1 - 2x}, reflecting the two choices (0 or 1) at each position. More generally, in the ring of formal power series, the geometric series enables algebraic manipulations like inversion, where \frac{1}{1 - r x} represents exponential growth in coefficient extraction for combinatorial recurrences. Exponential generating functions (EGFs) adapt the geometric series for labeled structures, incorporating factorial scaling to account for permutations of labels. The EGF for the sequence b_n = n! (counting permutations of n labeled elements) is \sum_{n=0}^\infty n! \frac{x^n}{n!} = \sum_{n=0}^\infty x^n = \frac{1}{1 - x}, a geometric series that captures the exponential growth in labeled enumerations. This form extends to other labeled objects, such as cycles or matchings, where the geometric series provides a baseline for exponential growth adjusted by combinatorial factors. In formal power series contexts, EGFs with geometric terms facilitate analysis of structures like trees or graphs, emphasizing the role of labeling in scaling coefficients. A key example of geometric series integration appears in the OGF for the Fibonacci sequence, defined by F_0 = 0, F_1 = 1, and F_n = F_{n-1} + F_{n-2} for n \geq 2. The OGF is \sum_{n=0}^\infty F_n x^n = \frac{x}{1 - x - x^2}, which decomposes via partial fractions into geometric components: \frac{x}{1 - x - x^2} = \frac{\phi x}{1 - \phi x} - \frac{\hat{\phi} x}{1 - \hat{\phi} x}, where \phi = \frac{1 + \sqrt{5}}{2} and \hat{\phi} = \frac{1 - \sqrt{5}}{2} are the golden ratio and its conjugate, yielding Binet's formula upon coefficient extraction. Similarly, the partition function p(n), counting integer partitions of n, has OGF \sum_{n=0}^\infty p(n) x^n = \prod_{k=1}^\infty \frac{1}{1 - x^k}, an infinite product of geometric series where each factor \frac{1}{1 - x^k} = \sum_{m=0}^\infty x^{k m} enumerates multiples of part size k. Generating functions leveraging geometric series support algebraic operations that mirror combinatorial compositions. Multiplication of two OGFs or EGFs corresponds to the Cauchy convolution of their coefficients, combining sequences additively (e.g., disjoint unions of structures), while division inverts this to yield deconvolution, solving for constituent parts in recursive counts. These manipulations, rooted in the formal power series ring, enable systematic enumeration without explicit summation, as seen in deriving closed forms for sequences like binary strings or permutations from their geometric bases.

Applications

Finance and economics

In finance, geometric series are fundamental to modeling compound interest, where the future value of a principal amount P invested at an annual interest rate r compounded n times is given by A = P(1 + r)^n. This formula arises from the geometric sequence of balances, starting with P, then P(1 + r), P(1 + r)^2, up to P(1 + r)^n, representing the exponential growth due to interest earned on both principal and prior interest./06%3A_Money_Management/6.04%3A__Compound_Interest) The total amount can be viewed as the sum of a finite geometric series when considering the incremental interest contributions, though the closed-form expression directly captures the compounded effect. Perpetuities, representing infinite streams of constant cash flows C discounted at rate r, have a present value of PV = \frac{C}{r}, derived from the infinite geometric series \sum_{k=1}^{\infty} \frac{C}{(1 + r)^k} = \frac{C}{r} for | \frac{1}{1 + r} | < 1. This converges when r > 0, providing a key tool for valuing assets with unending payments, such as certain endowments or preferred stocks. Annuities extend this to finite durations, where the present value of n periodic payments C is PV = C \frac{1 - (1 + r)^{-n}}{r}, the sum of a finite geometric series \sum_{k=1}^{n} \frac{C}{(1 + r)^k}. This formula is widely used for loan amortizations and , adjusting for the over a limited horizon. In broader economic models, geometric series underpin inflation adjustments and growth projections; for instance, the future under constant rate i follows P_t = P_0 (1 + i)^t, a that scales nominal values to real terms across periods. Similarly, GDP growth at a steady rate g is modeled as GDP_t = GDP_0 (1 + g)^t, enabling forecasts of long-term economic expansion via the series sum for cumulative output. A practical application is valuing perpetual bonds, such as consols, where the price S of a bond paying annual coupon C at yield i is S = \frac{C}{i}, obtained by summing the infinite discounted coupons as a geometric series \sum_{k=1}^{\infty} \frac{C}{(1 + i)^k}. For example, a bond with C = 50 and i = 0.05 yields S = 1000, illustrating how convergence ensures finite valuation despite infinite payments.

Physics and engineering

In physics, geometric series frequently model processes involving successive reductions or attenuations, such as . When the remaining mass of a radioactive substance is measured at , equal time intervals, the amounts form a geometric with a common r = 1 - \lambda, where \lambda is the per interval, approximating the continuous m(t) = m_0 e^{-\lambda t}. For example, if half the atoms decay every fixed period, r = 1/2, and the total decayed over n steps is the partial sum of the series, converging to m_0 as n \to \infty for |r| < 1. This model is useful in simulations and half-life calculations. Zeno's paradoxes, particularly the dichotomy paradox, illustrate the convergence of geometric series in resolving apparent infinities in motion. To traverse a finite distance d, one must first cover d/2, then d/4, then d/8, and so on, forming an infinite geometric series with first term a = d/2 and ratio r = 1/2. The sum is s = a / (1 - r) = d, showing the total distance (and similarly time, assuming constant speed) is finite despite infinitely many steps. This mathematical resolution, rooted in the properties of convergent series with |r| < 1, reconciles the paradox by demonstrating that infinite subdivisions do not imply infinite extent. In electrical engineering, geometric series approximate the behavior of RC circuits in discrete-time analyses, such as numerical simulations. The continuous charging voltage across the capacitor is V(t) = V_0 (1 - e^{-t / \tau}), where \tau = RC, but in discrete steps using methods like forward Euler, the voltage updates as V_{n+1} = V_n + (V_0 - V_n) \Delta t / \tau = V_n (1 - \Delta t / \tau) + V_0 \Delta t / \tau, yielding a geometric sequence for the transient term with ratio r = 1 - \Delta t / \tau < 1. This approximation converges to the exact exponential as \Delta t \to 0, aiding in digital circuit design and signal processing. Wave reflections in optics and acoustics often sum as infinite geometric series to determine overall transmission. In a Fabry-Pérot interferometer, light undergoes multiple bounces between two partially reflective mirrors, with the transmitted amplitude being the sum E_t = E_i t [1 + r^2 e^{i \phi} + (r^2 e^{i \phi})^2 + \cdots], where t is the single-pass transmission, r the reflection coefficient (|r| < 1), and \phi = 4\pi L / \lambda the phase shift for cavity length L. The series sums to E_t = E_i t / (1 - r^2 e^{i \phi}), yielding the transmission coefficient T = |E_t / E_i|^2 = 1 / (1 + F \sin^2(\phi/2)), where F = 4R / (1 - R)^2 and R = |r|^2; peaks occur at resonant wavelengths. Similar series model acoustic wave transmission through layered media, summing infinite reflections for effective coefficients. Fractals like the Koch snowflake employ geometric series to quantify infinite yet bounded structures. Starting from an equilateral triangle of side length s and area A_0 = \sqrt{3} s^2 / 4, each iteration adds protrusions, increasing the perimeter by a factor of $4/3: the first-stage perimeter is $3s, second $3s \cdot (4/3), and nth P_n = 3s (4/3)^n, forming a divergent \sum (4/3)^k with |r| = 4/3 > 1, so the total perimeter is infinite. Conversely, the added area per stage is a with ratio $1/3: stage 1 adds $3 \cdot (\sqrt{3}/4) (s/3)^2 = A_0 / 9, stage 2 adds $12 \cdot A_0 / 81 = 4 A_0 / 81, and so on, summing to A = A_0 (1 + 8/5) = (8/5) A_0, finite despite infinite iterations. This highlights how distinguish divergent perimeters from convergent areas in self-similar fractals.

Biology and population dynamics

In biology, geometric series arise prominently in models of exponential population growth, particularly in discrete-time frameworks where populations reproduce at constant rates without environmental constraints. The discrete model describes population size at time t as N_t = N_0 r^t, where N_0 is the initial population and r is the growth ratio (equal to the net reproductive rate, often r = 1 + R with R as the per capita growth rate). The total population accumulated over n+1 time steps forms a finite geometric series: \sum_{t=0}^{n} N_t = N_0 \frac{r^{n+1} - 1}{r - 1} for r \neq 1, illustrating how unchecked growth leads to rapid proliferation. A classic example is bacterial population dynamics through binary fission, where each divides into two, yielding a growth ratio of r = 2. Starting from a single (N_0 = 1), the population after n generations is $2^n, but the cumulative number of cells produced in a reproduction tree—accounting for all divisions across generations—sums to a geometric series: N = 2^n - 1, representing the total offspring from the initial (e.g., 1 + 2 + 4 + ... + 2^{n-1}). This model applies to ideal lab conditions with unlimited resources, such as Escherichia coli cultures doubling every 20–30 minutes. In epidemic modeling, approximate the early phases of infectious disease spread within the Susceptible-Infectious-Recovered () framework, where the number of infectives grows exponentially when the R_0 > 1 and susceptibles are abundant. Discretely, this manifests as a I_t \approx I_0 R_0^t, with the series sum capturing cumulative cases before saturation effects like immunity or depletion of susceptibles introduce logistic adjustments. For instance, early outbreaks followed this pattern until behavioral changes and vaccinations altered transmission. When growth ratios exceed 1 without resource limits, geometric models predict divergence, culminating in a Malthusian catastrophe—where population overshoots , triggering , , or as a natural check. Thomas Malthus first described this in , noting populations increase geometrically while food supplies grow arithmetically, a principle echoed in modern for like rabbits in unbounded habitats. Fibonacci-like models of rabbit populations, originally posed by in 1202, generate sequences where each term is the sum of the prior two (e.g., breeding pairs: 1, 1, 2, 3, 5, ...), but the ratios of consecutive terms approach the \phi \approx 1.618, mimicking geometric growth in the long term. This age-structured approach refines pure geometric models by incorporating maturation delays, yet still relies on geometric series for asymptotic behavior in idealized, immortal populations. In bounded environments, such as those with density-dependent factors, geometric growth converges to an , preventing indefinite .

Computer science and algorithms

In , geometric series often provide closed-form solutions to optimize iterative computations, particularly in analyses. For instance, when implementing a for- to compute the sum of a geometric progression, such as S = \sum_{i=0}^{n-1} r^i where |r| < 1, directly iterating incurs O(n) time complexity. However, substituting the closed-form formula S = \frac{1 - r^n}{1 - r} eliminates the entirely, achieving O(1) time and avoiding unnecessary computations for large n. This optimization is routinely applied in algorithm design to enhance efficiency in numerical simulations and data processing pipelines. Divide-and-conquer algorithms frequently yield recurrences solvable via geometric series, revealing logarithmic complexities. Consider the recurrence T(n) = T(n/2) + 1 with T(1) = 1, modeling the time for algorithms like or 's divide step. Unfolding the recurrence produces T(n) = 1 + 1 + \cdots + 1 (log n times), yielding T(n) = O(\log n). This summation approach, rooted in , underpins the for broader divide-and-conquer recurrences, enabling precise big-O bounds without simulation. In hashing with separate chaining, geometric series model collision probabilities and chain lengths. Under simple uniform hashing, the load factor \alpha = n/m (where n is the number of keys and m the table size) determines expected chain behavior; the probability of examining at least k elements during an unsuccessful search follows a geometric distribution with success probability $1 - \alpha, leading to expected comparisons \sum_{k=1}^{\infty} k \alpha^{k-1} (1 - \alpha) = \frac{1}{1 - \alpha} = O(1 + \alpha). This ensures average O(1) operations when \alpha is constant, justifying chaining's efficiency over open addressing for high collision rates. Machine learning leverages geometric series to analyze optimization dynamics. In gradient descent for convex objectives, convergence rates are often geometric, with error reducing by a factor \kappa < 1 per iteration, such that \| \theta_t - \theta^* \| \leq \kappa^t \| \theta_0 - \theta^* \|, where \kappa depends on the condition number of the Hessian; this holds globally for strongly convex quadratics under appropriate step sizes. Similarly, in neural network backpropagation, gradient propagation through deep layers can be approximated using geometric series when weights yield spectral radius less than 1, as in \sum_{k=0}^{\infty} C^k = (I - C)^{-1} for Jacobian C, mitigating vanishing gradients in residual architectures. Amortized analysis in big-O notation frequently employs geometric series for dynamic data structures like resizable arrays. When appending elements and resizing by doubling capacity (from $2^i to $2^{i+1}), the total copying cost over n insertions is \sum_{i=0}^{\log n} 2^i = 2^{ \log n + 1 } - 1 = O(n), yielding O(1) amortized time per operation via the aggregate method. This geometric growth prevents linear resizing pitfalls, ensuring scalability in languages like C++'s std::vector.

Advanced Topics and History

Generalizations beyond real and complex numbers

Geometric series can be generalized to settings beyond the real and complex numbers, where traditional notions of convergence may not apply or require alternative topologies. In the context of formal power series, the series is treated algebraically without regard to convergence. A formal power series over a commutative ring R is an expression of the form \sum_{k=0}^\infty a_k x^k with coefficients a_k \in R, and the set of all such series forms the ring R[]. In this ring, the geometric series \sum_{k=0}^\infty x^k is formally summed to \frac{1}{1-x} for x \neq 1, as multiplication by $1-x yields 1, establishing the inverse without analytic considerations. This algebraic structure is fundamental in commutative algebra and enumerative combinatorics, where formal sums encode generating functions for combinatorial objects. In non-Archimedean fields, equipped with a valuation-based |\cdot| satisfying the ultrametric inequality, convergence of geometric series \sum_{k=0}^\infty a r^k occurs when |r| < 1, analogous to the real case but with disks of convergence defined by the valuation rather than the Euclidean metric. The formula remains \frac{a}{1-r} within this radius, but the topology allows for "inverted" convergence behaviors, such as series diverging for |r| > 1 while converging for |r| \leq 1 in certain extensions. p-adic numbers \mathbb{Q}_p provide a example: for a prime p, the p-adic | \cdot |_p ensures that \sum_{k=0}^\infty r^k converges to \frac{1}{1-r} if |r|_p < 1, with partial s approaching the limit in the p-adic metric. For instance, in the 2-adic numbers, the series \sum_{k=0}^\infty 2^k converges to -1 since |2|_2 = 1/2 < 1. This framework extends to general complete non-Archimedean fields, where rings converge uniformly on valuation balls. q-series represent another generalization, replacing the common ratio r with a parameter q (typically $0 < |q| < 1) to deform classical series into q-analogs, often linked to partition theory and basic hypergeometric functions. A basic q-geometric series takes the form \sum_{k=0}^\infty \frac{(a; q)_k}{(q; q)_k} z^k = {}_1\phi_0(a; -; q, z), where (b; q)_k = \prod_{j=0}^{k-1} (1 - b q^j) is the , converging for appropriate |z| in the complex or p-adic sense. This extends the ordinary geometric series \sum_{k=0}^\infty z^k = \frac{1}{1-z} (corresponding to a \to 0) and relates to infinite products like the \prod_{k=0}^\infty (1 - q^k z), with applications in quantum algebra and . Seminal work on these series emphasizes their role in q-analogs of expansions and integrals. Multivariable geometric series extend the univariate case to sums over multi-indices or lattices, such as \sum_{k_1=0}^\infty \cdots \sum_{k_d=0}^\infty r_1^{k_1} \cdots r_d^{k_d} = \prod_{i=1}^d \frac{1}{1 - r_i} for |r_i| < 1 in suitable topologies, including formal multivariable rings R[[x_1, \dots, x_d]]. Over lattices \mathbb{Z}^d, such series appear in the form \sum_{\mathbf{n} \in \mathbb{Z}^d \setminus \{\mathbf{0}\}} \prod_{i=1}^d r_i^{n_i} (with adjustments for ), connecting to multivariable functions like the Epstein zeta function \zeta(s; A) = \sum_{\mathbf{n} \in \mathbb{Z}^d \setminus \{\mathbf{0}\}} \frac{1}{(A \mathbf{n} \cdot \mathbf{n})^{s/2}}, which generalize the Riemann zeta via forms and arise in and physics for lattice sums. In non-Archimedean settings, convergence depends on the joint valuation of the ratios.

Historical development

The concept of the geometric series traces its roots to mathematics, where paradoxes involving infinite sums emerged in the 5th century BCE through the work of . , such as the and Achilles and the , implicitly challenged the notion of summing infinitely many terms by dividing distances into ever-smaller segments, raising foundational questions about that foreshadowed later developments in infinite series. In the 3rd century BCE, advanced these ideas through his , which approximated areas and volumes by inscribing and circumscribing polygons, effectively employing finite geometric series to bound curved figures like the parabola; his , for instance, summed a series of triangular areas to achieve precise results. During the medieval period, European and Indian mathematicians independently explored progressions and ratios that built upon ancient foundations. In 14th-century Europe, , a French scholar, developed graphical representations of varying qualities and ratios in his work on the latitude of forms, using coordinate-like methods to visualize proportional changes and investigate infinite series, including comparisons that anticipated . Concurrently, in 12th-century , Bhaskara II, in his treatise Lilavati, provided formulas for the sums of finite arithmetic and geometric progressions, applying them to practical problems in arithmetic and laying groundwork for systematic series summation in non-European traditions. The saw geometric series integrated into algebraic problem-solving, particularly for extracting roots of equations. Gerolamo Cardano, in his 1545 Ars Magna, employed radical expressions to provide general solutions to cubic and quartic equations. François Viète extended this by introducing symbolic notation in works like Zeteticae Logisticae (late ), framing algebraic relations geometrically to solve higher-degree polynomials, thereby elevating algebraic methods within a systematic analytic framework. In the 17th and 18th centuries, the development of intertwined geometric series with foundational analysis. and independently incorporated infinite geometric series into their fluxion and differential methods, using them to represent and compute integrals, as seen in Newton's expansions for arbitrary exponents. Leonhard Euler advanced this further in the mid-18th century, deriving decompositions—such as for the sine —by factoring into geometric series, which enabled profound insights into and theory. In the modern era, geometric series found new applications in and non-Archimedean analysis. Bernhard Riemann's 1859 investigation of the zeta function extended Euler's product formula, employing geometric series regularization to analytically continue the function beyond its original domain, providing a rigorous framework for summing in number-theoretic contexts. At the turn of the , Kurt introduced p-adic numbers around 1900, where geometric series with |r|_p < 1 converge in this ultrametric , enabling solutions to equations modulo primes that lift uniquely and revolutionizing .

References

  1. [1]
    Calculus II - Special Series - Pauls Online Math Notes
    Aug 13, 2024 · ... series term are numbers raised to a power. This means that it can be put into the form of a geometric series. We will just need to decide ...
  2. [2]
    CC Geometric Series
    A finite geometric series S n is a sum of the form , (7.3) where a and r are real numbers such that . r ≠ 1.
  3. [3]
    [PDF] Lecture 16: Geometric series - Harvard Mathematics Department
    16.5. Historically, the geometric series first appeared in Zeno's paradox. Assume. Archilles races a turtle who is given an advance of a = 1 miles Assume the ...
  4. [4]
    Geometric Sequences - Student Academic Success
    Geometric sequence. s are used in a wide range of real-world applications due to their ability to model exponential changes. They may be used in the ...
  5. [5]
    7.3 - Geometric Sequences
    A geometric sequence is a sequence in which the ratio consecutive terms is constant. Common Ratio. Since this ratio is common to all consecutive pairs of terms, ...
  6. [6]
  7. [7]
    Infinite geometric series word problem: repeating decimal (video)
    Aug 8, 2016 · According to Sal's method, any repeating decimal can be expressed as an infinite geometric series with r = 0.1 or 0.01 or 0.001 or 0.0001 or so on.Missing: finite | Show results with:finite
  8. [8]
    Worked example: divergent geometric series (video) - Khan Academy
    Jun 28, 2018 · Sal evaluates the infinite geometric series -0.5+1.5-4.5+... Because the common ratio's absolute value is greater than 1, the series doesn't converge.
  9. [9]
    [PDF] Deriving the Formula for the Sum of a Geometric Series - UMD MATH
    x x xx x . Finally, dividing through by 1 – x, we obtain the classic formula for the sum of a geometric series: x.
  10. [10]
    2.4 Finite Geometric Series
    is given by a simple formula, which we will now derive. 2.22 Theorem (Finite Geometric Series.) Let $r$ be a real number such that $r\neq 1$ . Then for all ...
  11. [11]
    [PDF] Math 2300: Calculus II Geometric series Goal
    A geometric sequence is defined by an initial term a and constant ratio r, and looks like this: a, ar, ar2, ar3, .... Partial geometric sums are just the partial ...
  12. [12]
    [PDF] Lecture 18: Geometric series - Nathan Pflueger
    Oct 19, 2011 · Geometric series are the main example of infinite series. They provide an important class of examples for the.
  13. [13]
    [PDF] Summations - OSU Math
    The value of a finite series is always well defined, and its terms can be added in any order. Given a sequence a1, a2, of numbers, the infinite sum a₁ + a₂+ can ...<|control11|><|separator|>
  14. [14]
    [PDF] Sequences and Series - UCLA Math Circle
    Oct 29, 2023 · A finite geometric series is one where the number of values in the sequence is known to us, and thus we can add them all up to get our sum.
  15. [15]
    [PDF] Series
    The geometric series formulas allow us to eval- uate any series whose terms involve only exponential functions like 2n or 2−n, but not power functions like n or ...<|control11|><|separator|>
  16. [16]
    [PDF] 5 Mathematical Induction - Computer Science
    Apr 5, 2021 · PROOFS BY MATHEMATICAL INDUCTION. 511. (Thus the sequence is hx0, x0 · α, x0 · α. 2, x0 · α. 3,...i.) A geometric series or geometric sum is ∑.
  17. [17]
    [PDF] Induction, Recursion, and Recurrences - Dartmouth Computer Science
    A finite geometric series with common ratio r is a sum of the form "n−1 i=0 ... is straightforward to prove by induction, or with the formula for the sum of a ...
  18. [18]
    [PDF] INFINITE SERIES 1. Introduction The two basic concepts of calculus ...
    If we start summing a geometric series not at 1, but at a higher power of x, then we can still get a simple closed formula for the series, as follows.
  19. [19]
    [PDF] Math 141: Lecture 18 - Sequences and infinite series
    Nov 14, 2016 · Since the odd partial sums are decreasing and bounded below, they converge to a limit o. Since the even partial sums are increasing and bounded ...
  20. [20]
    [PDF] Geometric series theorem
    If |r| ≥ 1, the geometric series diverges. Proof: We start from ... +rn) = (1−r)·1+(1−r)r+(1−r)r2+ ... +(1−r)rn = 1−r+r−r2+r2−r3+ ... +rn−rn+1 ...
  21. [21]
    [PDF] Infinite Series
    Here are examples of convergence, divergence, and oscillation: The first series converges. Its next term is 118, after that is 1116-and every step brings us ...Missing: criteria | Show results with:criteria
  22. [22]
    [PDF] 1 Basics of Series and Complex Numbers
    1.3 Geometric sums and series. For any complex number q 6= 1, the geometric sum. 1 + q + q2 + ··· + qn = 1 − qn+1. 1 − q . (10). To prove this, let Sn =1+ q + ...
  23. [23]
    Remainders for Geometric and Telescoping Series - Ximera
    For a convergent geometric series or telescoping series, we can find the exact error made when approximating the infinite series using the sequence of partial ...
  24. [24]
    Calculus II - Estimating the Value of a Series
    Nov 16, 2022 · So, the remainder tells us the difference, or error, between the exact value of the series and the value of the partial sum that we are using as ...
  25. [25]
    [PDF] Infinite Series of Complex Numbers - Trinity University
    A series Pak is called absolutely convergent provided P|ak| converges. Note that there is no a priori connection between convergence and.
  26. [26]
    Calculus II - Power Series and Functions - Pauls Online Math Notes
    Nov 16, 2022 · In this section we discuss how the formula for a convergent Geometric Series can be used to represent some functions as power series.
  27. [27]
    Calculus II - Ratio Test - Pauls Online Math Notes
    Aug 13, 2024 · In this section we will discuss using the Ratio Test to determine if an infinite series converges absolutely or diverges. The Ratio Test can ...
  28. [28]
    How Isaac Newton Discovered the Binomial Power Series
    Aug 31, 2022 · He recounts how he made a major discovery equating areas under curves with infinite sums by a process of guessing and checking.
  29. [29]
    Calculus II - Taylor Series - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will discuss how to find the Taylor/Maclaurin Series for a function. This will work for a much wider variety of function ...
  30. [30]
    [PDF] generatingfunctionology - Penn Math
    May 21, 1992 · This book is about generating functions and some of their uses in discrete mathematics. The subject is so vast that I have not attempted to give ...
  31. [31]
    [PDF] Enumerative Combinatorics Volume 1 second edition - Mathematics
    Enumerative combinatorics is about counting elements of a finite set. This book is an updated second edition with new sections and exercises.
  32. [32]
    Fibonacci Number -- from Wolfram MathWorld
    ### Generating Function for Fibonacci Numbers and Relation to Geometric Series
  33. [33]
    Partition Function P -- from Wolfram MathWorld
    ### Generating Function Formula for Partition Function P and Relation to Geometric Series
  34. [34]
    [PDF] The Mathematics of Money - University of Kentucky
    Here is a formula for the sum of the first N terms in a geometric sequence (in the case that c 6= 1):. P + cP + cP2 + ··· + cN−1P = P. cN − 1 c − 1 .
  35. [35]
    Perpetuities - NetMBA
    Because this cash flow continues forever, the present value is given by an infinite series: PV = C / ( 1 + i ) + C / ( 1 + i )2 + C / ( 1 + i ) ...
  36. [36]
    [PDF] Geometric Annuities
    Geometric Perpetuity: If the annuity is a perpetuity, then the geometric sum representing its present value is actually a convergent geometric series with.
  37. [37]
    Annuities | CK-12 Foundation
    The future value (FV) of an annuity can be calculated using a geometric series formula: F V = R ∗ ( ( 1 + i ) n − 1 ) i , where R is the payment, i is the ...
  38. [38]
    [PDF] Geometric Series and Annuities - UCR Math Department
    Our goal here is to calculate annuities. For example, how much money do you need to have saved for retirement so that you can withdraw a fixed.
  39. [39]
    Sequences and Series in Economics: Understanding Long-Term ...
    Geometric sequences are frequently used in economics to model exponential growth or decay, such as investment returns, population growth, or inflation. The ...
  40. [40]
    [PDF] 3. VALUATION OF BONDS AND STOCK Investors Corporation
    A perpetual bond has face value $1,000, and coupon 8%. You bought this bond ... This is an infinite geometric series with first term a = D1. 1 + R. , and ...
  41. [41]
    [PDF] Geometric Series, Perpetuity and Term Annuity
    perpetuity is: P perp, T. $A r. (1 r)T. $A. 1 r (1 r)T. Subtracting the present value of a perpetuity that has first payment at T+1 from a perpetuity valued.
  42. [42]
    [PDF] The `radioactive dice' experiment: why is the `half-life' slightly wrong?
    So we see that the decay is given by a geometric progression. The number of undecayed dice obtained using this method for the first 12 throws is shown in table ...Missing: sequence | Show results with:sequence
  43. [43]
    Exponential Decay - an overview | ScienceDirect Topics
    Exponential decay occurs when the growth rate is negative. In the case of a discrete domain of definition with equal intervals, it is also called geometric ...
  44. [44]
    [PDF] Zeno's Paradox and Geometric Series - Cornell Mathematics
    Dichotomy paradox. We assume that a finite distance can be traversed in a finite time - for example an arrow being fired at a target. To traverse the ...
  45. [45]
    Interferometers and the Fabry-Perot Interferometer - Benjamin Klein
    The overall transmission of the interferometer is therefore proportional to the plot of intensity inside the cavity, above. The Fabry-Perot resonator acts like ...
  46. [46]
    [PDF] GEOMETRIC AND EXPONENTIAL POPULATION MODELS
    Geometric models are discrete-time, while exponential models are continuous-time. Both are related to population dynamics, and the geometric model behaves ...
  47. [47]
    Bacterial Population Growth (Theory) : Ecology Virtual Lab
    ... increase in numbers according to a geometric series. Exponential growth occurs in relation with the population increases by a certain factor with respect to ...
  48. [48]
    Module 5 : MICROBIAL GROWTH AND CONTROL - NPTEL Archive
    In the laboratory, under favorable conditions, a growing bacterial population doubles at regular intervals. Growth is by geometric progression: 1, 2, 4, 8, etc.Missing: series | Show results with:series<|separator|>
  49. [49]
    Mathematical models to characterize early epidemic growth: A Review
    The SIR model and derivatives is the framework of choice to capture population-level processes. The basic SIR model, like many other epidemiological models, ...
  50. [50]
    MATHEMATICA TUTORIAL, Part 1.2: Populations
    Thomas Malthus (1766--1834) was one of the first to note that populations grow with a geometric pattern (dP/dt = r P) while contemplating the fate of humankind.<|separator|>
  51. [51]
    17.2D: Malthus' Theory of Population Growth - Social Sci LibreTexts
    Feb 19, 2021 · Malthusian catastrophes refer to naturally occurring checks on population growth such as famine, disease, or war. These Malthusian catastrophes ...Missing: biology | Show results with:biology
  52. [52]
    2.1: Fibonacci's Rabbits - Mathematics LibreTexts
    Jul 17, 2022 · The growth of Fibonacci's rabbit population is presented in Table 2.1. At the start of each month, the number of juvenile, adult, and total number of rabbits ...Missing: geometric | Show results with:geometric
  53. [53]
    [PDF] Lecture 3: Summations and Analyzing Programs with Loops
    Feb 3, 1998 · To analyze program loops, write them as summations, working from inside out, and then solve the summations.
  54. [54]
    [PDF] Solving Divide-and-Conquer Recurrences
    A divide-and-conquer algorithm consists of three steps: • dividing a problem into smaller subproblems. • solving (recursively) each subproblem.
  55. [55]
    [PDF] Lecture 4: Hashing, Chaining, and Probing Analysis - Rice University
    Sep 10, 2024 · Chaining is a technique used to handle collisions in hashmaps. Because there is the potential that two different keys are hashed to the same ...
  56. [56]
    [PDF] Fast global convergence of gradient methods for high-dimensional ...
    projected gradient descent (3) enjoys a globally geometric convergence rate, meaning that there is some κ ∈ (0,1) such that1 kθt − bθk2 % κt kθ0 − bθk2.
  57. [57]
    [PDF] a solution to the vanishing/exploding gradient of deep neural networks
    Oct 3, 2022 · The formal geometric series P. ∞ k=0 Ck = (I − C)−1 holds for all squared matrices C having spectral radius less than 1. Therefore, by ...
  58. [58]
    [PDF] 1 The ubiquitous example: Dynamic arrays (lists)
    Sep 5, 2022 · In this lecture we discuss a useful form of analysis, called amortized analysis, for problems in which one must perform a series of ...
  59. [59]
    [PDF] Formal power series
    Nov 7, 2012 · Formal power series are like polynomials with infinite terms, where the focus is on the sequence of coefficients, not convergence.
  60. [60]
    [PDF] Enumerative Combinatorics 2: Formal power series
    Nov 4, 2013 · A formal power series over R is an infinite sequence (r0, r1, r2, ...) represented as r0 + r1x + r2x2 + ... = X rnxn.
  61. [61]
    [PDF] Infinite series in p-adic fields - Keith Conrad
    At x = 1 the series does not converge, so the geometric series has disc of convergence {x ∈ K : |x| < 1}, which we already knew. In our next three examples ...
  62. [62]
    [PDF] Lecture Notes For An Introductory Minicourse on q-Series
    Sep 19, 1995 · We start by considering q-extensions (also called q-analogues) of the binomial theorem, the exponential and gamma functions, and of the beta ...
  63. [63]
    [PDF] Nicole Oresme and the Revival of Medieval Mathematics
    This paper will attempt to demonstrate this by looking at Oresme's contributions to mathematics through his studies of infinite series and his development of ...
  64. [64]
    [PDF] 9. Mathematics in the sixteenth century - UCR Math Department
    . The following anecdote regarding Viète illustrates his use of trigonometry to find roots of polynomials. In 1593, A. van Roomen (1561 – 1615) issued an ...
  65. [65]
    [PDF] classics of mathematics
    For his pioneering work on symbolism Viète is often called the father of modern algebraic notation. Viète made other contributions to algebra. In 1615, he ...
  66. [66]
    [PDF] 12. The development of calculus 13. Newton and Leibniz
    In the early 17th century mathematicians became interested in several types of problems, partly because these were motivated by advances in physics and partly ...
  67. [67]
    [PDF] How Euler discovered the zeta function
    Euler's idea was to start with the familiar formula for the sum of a geometric series: 1. 1 − x. =1+ x + x2 + x3 + ... (0 <x< 1). For any prime number p and ...
  68. [68]
    [PDF] the zeta function and the riemann hypothesis - UChicago Math
    Oct 27, 2014 · History. The zeta function, usually referred to as the Riemann zeta function today, has been studied in many different forms for centuries.
  69. [69]
    A course in p-adic analysis , by Alain M. Robert, Graduate Texts in ...
    Oct 11, 2001 · The p-adic numbers were discovered by K. Hensel around the end of the nine- teenth century. In the course of one hundred years, ...Missing: 20th | Show results with:20th