Fact-checked by Grok 2 weeks ago

Small-angle approximation

The small-angle approximation is a fundamental mathematical technique used in physics and to simplify calculations involving when the angle θ is small and measured in radians. For such angles, typically θ ≪ 1 (e.g., less than about 0.1 radians or 6°), the approximations sin θ ≈ θ, tan θ ≈ θ, and cos θ ≈ 1 hold with high accuracy, as derived from the first few terms of the expansions of these functions around θ = 0. These relations stem from the power series representations: sin θ = θ - θ³/6 + ..., cos θ = 1 - θ²/2 + ..., and tan θ = θ + θ³/3 + ..., where higher-order terms become negligible for small θ. The approximation introduces errors on the order of θ³ or smaller, often less than 0.2% for θ < 0.1 radians, making it practical for many analytical solutions. This approximation originates from the limiting behavior of trigonometric functions as θ approaches zero, where the derivative of sin θ at 0 is cos 0 = 1, yielding the tangent line y = θ as a linear approximation. It is particularly valuable in contexts where exact solutions are intractable, such as deriving the simple harmonic motion equation for , where the restoring torque is linearized to sin θ ≈ θ, leading to a period independent of amplitude for small oscillations. In optics, it underpins the for ray tracing in lenses and mirrors, assuming rays are close to the optical axis with small angles to enable linear equations. Beyond these, the small-angle approximation facilitates computations in astronomy for estimating object sizes from angular diameters, using θ ≈ D/d where D is the physical diameter and d the distance, especially when d ≫ D. It also appears in wave mechanics, vector analysis, and numerical simulations, where it reduces nonlinear problems to linear ones for faster convergence and insight. While powerful, its validity diminishes for larger angles, necessitating full trigonometric evaluations or higher-order expansions in precise modeling.

Fundamentals

Definition and Basic Approximations

The small-angle approximation refers to the simplification of trigonometric functions and related expressions when the angle θ is much smaller than 1 radian, typically satisfying θ ≪ 1, allowing higher-order terms in their series expansions to be neglected for practical computations. This approach is particularly useful in and where exact trigonometric evaluations are cumbersome, providing a linear or quadratic estimate that maintains sufficient accuracy for small deviations from zero. The primary approximations for the basic trigonometric functions are derived from their Taylor series expansions around θ = 0. Specifically, for θ in radians, sin θ ≈ θ, cos θ ≈ 1 - (θ²/2), and tan θ ≈ θ. These relations hold because the leading terms dominate when θ is small, with the sine and tangent functions approximating the angle itself linearly, while cosine deviates quadratically from unity. From these, derived approximations for the reciprocal functions follow directly: sec θ ≈ 1 + (θ²/2), csc θ ≈ 1/θ, and cot θ ≈ 1/θ - (θ/3). The use of radians is essential for these approximations, as the Taylor series coefficients are dimensionless only in this unit system, ensuring the approximations scale correctly without additional conversion factors. These approximations trace their origins to early 18th-century developments in calculus, notably the infinite series expansions introduced by in his 1715 work Methodus Incrementorum Directa et Inversa and specialized by in his 1742 treatise , which formalized the basis for truncating series at low orders for small arguments.

Assumptions and Units

The small-angle approximation relies on the fundamental assumption that the angle θ is sufficiently small relative to the scale of the problem, typically in the limit as θ approaches zero, where higher-order terms in the series expansion become negligible. This condition ensures that perturbations or deviations from the idealized small-angle regime do not significantly affect the accuracy, as the approximation is asymptotic in nature—its precision improves monotonically as θ decreases toward zero. A critical requirement for the numerical validity of these approximations is that angles must be expressed in radians, the natural unit for trigonometric functions derived from their Taylor series expansions around θ = 0. Using degrees without conversion introduces substantial errors because the series coefficients are calibrated specifically for the radian measure, where the arc length equals the radius for θ = 1; in degrees, the equivalent "small" angle would require rescaling by π/180, rendering the direct substitution sin θ ≈ θ invalid. For reference, 1 radian is approximately 57.3 degrees, highlighting why unadjusted degree-based approximations fail to capture the linear behavior near zero. The approximation sin θ ≈ θ has a relative error of approximately 1% at θ ≈ 0.25 radians (about 14°), with error increasing for larger angles. For example, at θ = 0.1 radians (≈ 5.7°), the percentage error is approximately 0.17%, calculated as |(sin 0.1 - 0.1)/sin 0.1| × 100. At θ = 0.22 radians (≈ 12.6°), the error rises to about 0.81%, with sin(0.22) ≈ 0.2182 versus the approximation 0.22. These examples illustrate the rapid improvement in accuracy for smaller angles, underscoring the approximation's utility in contexts demanding high precision for θ ≪ 1 radian.

Justifications

Geometric Justification

The small-angle approximation for sine arises from considering a unit circle, where an angle \theta (in radians) subtends an arc of length \theta. In this setup, the chord length opposite the angle is \sin \theta, and as \theta approaches zero, the arc length and chord length become indistinguishable, leading to \sin \theta \approx \theta. This can be visualized by inscribing a sector with a small angle \theta at the center; the vertical rise along the ray equals the arc length in the limit, confirming the approximation geometrically without relying on series expansions. A more rigorous geometric justification uses the squeeze theorem on areas within the unit circle diagram. Consider three regions sharing the origin: a small triangle with area \frac{1}{2} \cos \theta \sin \theta, the circular sector (wedge) with area \frac{1}{2} \theta, and a larger triangle with area \frac{1}{2} \tan \theta. Since the small triangle is contained within the sector, which is contained within the large triangle, the area inequalities yield \cos \theta \leq \frac{\sin \theta}{\theta} \leq \frac{1}{\cos \theta} for $0 < \theta < \frac{\pi}{2}. As \theta \to 0, both bounds approach 1, squeezing \lim_{\theta \to 0} \frac{\sin \theta}{\theta} = 1, thus \sin \theta \approx \theta for small \theta. For cosine, the approximation \cos \theta \approx 1 follows from the same unit circle geometry. The point on the circle at angle \theta has x-coordinate \cos \theta, which represents the horizontal projection from the origin. As \theta shrinks toward zero, this projection approaches the full radius of 1, with the vertical offset \sin \theta becoming negligible; the "sagitta" (curved deviation from the straight line) vanishes proportionally to \theta^2, leaving \cos \theta arbitrarily close to 1. The tangent approximation \tan \theta \approx \theta emerges from a right triangle inscribed in the unit circle, where the opposite side over the adjacent side (of length 1) gives \tan \theta. For small \theta, the ray at angle \theta intersects the vertical line at x=1 (the tangent line to the circle at (1,0)) at a height approximately equal to the arc length \theta, since the circle and the tangent line coincide near the x-axis; thus, the opposite side approaches \theta. This is evident in the skinny triangle formed by the origin, the point (1,0), and the intersection point (1, \tan \theta), where the hypotenuse closely hugs the adjacent side as \theta \to 0. In skinny triangles—narrow right triangles with a small apex angle \theta and long equal sides—the hypotenuse approximates the adjacent side, reinforcing \sin \theta \approx \tan \theta \approx \theta. As \theta decreases, the differences between the arc length, chord (sine), and tangent segment diminish proportionally, illustrating the unified geometric limit where all three measures converge for infinitesimal angles.

Calculus-Based Justification

The small-angle approximation for trigonometric functions arises from the Taylor series expansions of these functions around θ = 0, where higher-order terms become negligible for small values of θ (typically in radians). The Taylor series, also known as the when expanded about zero, represents a function as an infinite sum of terms calculated from its derivatives at that point. For the sine function, the Maclaurin series is \sin \theta = \theta - \frac{\theta^3}{3!} + \frac{\theta^5}{5!} - \frac{\theta^7}{7!} + \cdots, where the series alternates signs and involves odd powers of θ. For small θ, terms involving θ³ and higher powers are much smaller than θ, so truncating after the first term yields \sin \theta \approx \theta. Similarly, the Maclaurin series for cosine is \cos \theta = 1 - \frac{\theta^2}{2!} + \frac{\theta^4}{4!} - \frac{\theta^6}{6!} + \cdots, featuring even powers of θ with alternating signs. Neglecting terms beyond the constant and quadratic yields \cos \theta \approx 1 - \frac{\theta^2}{2}, which is the leading approximation for small angles. This expansion justifies the small-angle behavior where cosine is close to unity, with a small quadratic correction. The tangent function's approximation follows from the ratio of sine to cosine. Substituting the series gives \tan \theta = \frac{\sin \theta}{\cos \theta} \approx \frac{\theta}{1 - \frac{\theta^2}{2}}. For small θ, the denominator can be expanded using the binomial approximation (1 - u)^{-1} \approx 1 + u where u = \frac{\theta^2}{2}, yielding \frac{\theta}{1 - \frac{\theta^2}{2}} \approx \theta \left(1 + \frac{\theta^2}{2}\right) \approx \theta + \frac{\theta^3}{2}. Truncating at the linear term provides \tan \theta \approx \theta, consistent with the first-order approximations for sine and cosine. The full for tangent is \tan \theta = \theta + \frac{1}{3}\theta^3 + \frac{2}{15}\theta^5 + \frac{17}{315}\theta^7 + \cdots. The validity of these truncations is supported by Taylor's theorem, which includes a remainder term estimating the error after n terms. In Lagrange's form, the remainder R_n(\theta) after the polynomial of degree n is R_n(\theta) = \frac{f^{(n+1)}(\xi)}{(n+1)!} \theta^{n+1}, where ξ is between 0 and θ. For sine approximated by the degree 1 polynomial θ (n=1), R_1(\theta) = \frac{f''(\xi)}{2!} \theta^2 = -\frac{\sin \xi}{2} \theta^2, bounded by \frac{\theta^2}{2} in absolute value since |\sin \xi| \leq 1, which diminishes as θ approaches zero. Similar bounds apply to cosine and tangent, confirming that the approximations improve as θ becomes small, with the error scaling with higher powers of θ. These series expansions were formalized by Brook Taylor in his 1715 work Methodus Incrementorum Directa et Inversa, introducing the general theorem, and further developed by Colin Maclaurin in his 1742 Treatise on Fluxions, which applied them systematically to functions like sine and cosine.

Algebraic Justification

The small-angle approximation for the sine function can be justified algebraically through the limit \lim_{\theta \to 0} \frac{\sin \theta}{\theta} = 1, which holds when \theta is measured in radians and implies \sin \theta \approx \theta for small \theta. This limit is established using the squeeze theorem applied to geometric inequalities derived from the unit circle, combined with basic trigonometric identities such as \sin \theta < \theta < \tan \theta for $0 < \theta < \pi/2. An intuitive algebraic reinforcement of this approximation involves iterative application of the double-angle identity \sin \theta = 2 \sin(\theta/2) \cos(\theta/2). For sufficiently small \theta, the half-angle \theta/2 is even smaller, where \cos(\theta/2) \approx 1 holds approximately, yielding \sin \theta \approx 2 \sin(\theta/2). Substituting the approximation recursively, \sin(\theta/2) \approx \theta/2, gives \sin \theta \approx 2 \cdot (\theta/2) = \theta. This process can be repeated by halving the angle multiple times, approaching the limit behavior algebraically without invoking infinite series. For the cosine function, the approximation \cos \theta \approx 1 - \theta^2/2 follows algebraically from the known sine limit and the identity $1 - \cos \theta = 2 \sin^2(\theta/2). Substituting into the rearranged limit form, \lim_{\theta \to 0} \frac{1 - \cos \theta}{\theta^2} = \frac{1}{2}, yields: \frac{1 - \cos \theta}{\theta^2} = \frac{2 \sin^2(\theta/2)}{\theta^2} = \frac{1}{2} \left( \frac{\sin(\theta/2)}{\theta/2} \right)^2. As \theta \to 0, \frac{\sin(\theta/2)}{\theta/2} \to 1, confirming the limit value of $1/2 and thus the approximation. The tangent approximation \tan \theta \approx \theta is derived algebraically by substituting the sine and cosine approximations into the definition \tan \theta = \sin \theta / \cos \theta: \tan \theta \approx \frac{\theta}{1 - \theta^2/2} \approx \theta \left(1 + \frac{\theta^2}{2}\right) \approx \theta, where the higher-order term \theta^3/2 is neglected for small \theta. This finite manipulation preserves the leading-order behavior without series expansion. Alternatively, these approximations can be obtained through finite polynomial fitting methods, such as least-squares approximation or interpolation, by matching polynomial coefficients to tabulated values of the trigonometric functions at small discrete points near \theta = 0. For instance, fitting a linear polynomial to \sin \theta over points in [-\epsilon, \epsilon] for small \epsilon > 0 yields the slope 1 and intercept 0, confirming \sin \theta \approx \theta; similar fits for \cos \theta and \tan \theta produce the quadratic and linear forms, respectively. These algebraic techniques rely on solving linear systems from the least-squares criterion \min \sum (f(\theta_i) - p(\theta_i))^2.

Error and Accuracy

Error Bounds

The absolute error in the small-angle approximation \sin \theta \approx \theta (with \theta in radians) satisfies |\sin \theta - \theta| \leq \frac{\theta^3}{6} for small positive \theta, as established by the alternating series estimation theorem applied to the of sine. The relative error for this approximation is then approximately \frac{\theta^2}{6}, since the true value \sin \theta is close to \theta for small angles. For the approximation \cos \theta \approx 1 - \frac{\theta^2}{2}, the absolute error is bounded by \frac{\theta^4}{24}, derived similarly from the remainder after the quadratic term. The corresponding relative error is approximately \frac{\theta^4}{24}, given that \cos \theta \approx 1. For \tan \theta \approx \theta, the leading error term from the is approximately \frac{\theta^3}{3}, yielding a relative error of roughly \frac{\theta^2}{3}. To illustrate these errors practically, consider the following table of relative errors (defined as |\frac{\text{approximation} - \text{true value}}{\text{true value}}| \times 100\%) for the \sin \theta \approx \theta approximation at selected small angles in radians:
\theta (radians)\theta (degrees)True \sin \thetaRelative Error (%)
0.1≈5.7°0.0998330.17
0.5≈28.6°0.4794264.29
1.0≈57.3°0.84147118.85
These values demonstrate rapid growth in error as \theta increases; for instance, the relative error remains below 1% for \theta \lesssim 0.2 radians (≈11.5°). In applications, the approximations are typically considered reliable to within 1% relative error for angles up to about 10° (≈0.175 radians), beyond which higher-order terms become significant and the linear or forms fail to capture the behavior accurately. For the \tan \theta \approx \theta case specifically, the 1% threshold occurs near 10°, while for \sin \theta \approx \theta it extends slightly further due to the smaller leading error term.

Higher-Order Terms

To achieve greater accuracy in the small-angle approximation for moderately small angles, higher-order terms from the expansions of the are incorporated. For the sine function, the approximation improves from \sin \theta \approx \theta to \sin \theta \approx \theta - \frac{\theta^3}{6}, where the cubic term corrects the leading-order error, leaving a remainder of O(\theta^5). Similarly, for cosine, the refinement is \cos \theta \approx 1 - \frac{\theta^2}{2} + \frac{\theta^4}{24}, with the quartic term reducing the error to O(\theta^6). For , the higher-order form is \tan \theta \approx \theta + \frac{\theta^3}{3}, truncating at the cubic term to achieve an error of O(\theta^5). These higher-order approximations are particularly useful for angles up to about 0.5 radians (approximately 28.6 degrees), where they substantially reduce the approximation error compared to the linear terms alone. The inclusion of the next significant term typically cuts the error by more than 90% for such angles, as the ratio of the leading neglected terms scales with \theta^2, which is less than 0.25 at \theta = 0.5. For example, at \theta = 0.3 radians, the basic approximation \sin \theta \approx \theta yields a relative error of approximately 1.5%, whereas the higher-order form \sin \theta \approx \theta - \frac{\theta^3}{6} reduces this to about 0.007%, demonstrating an error reduction exceeding 99%. The accuracy of these truncated expansions can be rigorously bounded using the Lagrange form of the remainder in . For a f(\theta) expanded around 0 to order n, the remainder after n terms is R_n(\theta) = \frac{f^{(n+1)}(\xi)}{(n+1)!} \theta^{n+1} for some \xi between 0 and \theta. For the sine approximation up to the cubic term (n=3), this becomes R_3(\theta) = \frac{\sin \xi}{24} \theta^4, where |\sin \xi| \leq 1, so |R_3(\theta)| \leq \frac{\theta^4}{24}. This bound provides a quantitative limit for practical use, though the true error scales as O(\theta^5) due to the structure of the sine series.

Extensions

Angle Sum and Difference Approximations

The small-angle approximation extends naturally to sums and differences of angles through the trigonometric formulas, particularly when one is small relative to the other or when both are small. Consider the sum of angles θ and δ, where δ is small (in radians). Using the formula, \sin(\theta + \delta) = \sin \theta \cos \delta + \cos \theta \sin \delta, and applying the basic approximations \cos \delta \approx 1 and \sin \delta \approx \delta, this simplifies to \sin(\theta + \delta) \approx \sin \theta \cdot 1 + \cos \theta \cdot \delta = \sin \theta + \delta \cos \theta. A similar derivation for the cosine function yields \cos(\theta + \delta) = \cos \theta \cos \delta - \sin \theta \sin \delta \approx \cos \theta - \delta \sin \theta. These forms are first-order approximations derived from the limit definitions underlying calculus, where the change in the function over a small increment δ approximates the derivative times δ. For the tangent function, the addition formula is \tan(\theta + \delta) = \frac{\tan \theta + \tan \delta}{1 - \tan \theta \tan \delta}. When both θ and δ are small, \tan \theta \approx \theta and \tan \delta \approx \delta, and the product \tan \theta \tan \delta is second-order small, so it can be neglected, leading to \tan(\theta + \delta) \approx \theta + \delta. More generally, if only δ is small, the approximation becomes \tan(\theta + \delta) \approx \tan \theta + \delta \sec^2 \theta, which aligns with the differential d(\tan \theta) = \sec^2 \theta \, d\theta. These extensions are useful in contexts requiring incremental changes, such as in differential equations or perturbation analysis. In applications involving differentials, the small-angle approximations manifest directly as d(\sin \theta) \approx \cos \theta \, d\theta and d(\cos \theta) \approx -\sin \theta \, d\theta, providing linear approximations for changes in angle. This is foundational in for deriving rates of change in trigonometric expressions. Regarding errors, when approximating sums or differences of independent small angles, the uncertainties such that the variance of the total angle is the sum of the individual variances, meaning errors add quadratically: if angles α and β have uncertainties Δα and Δβ, then the uncertainty in α ± β is \sqrt{(\Delta \alpha)^2 + (\Delta \beta)^2}. This follows from rules for and , ensuring the combined remains reliable when individual errors are small.

Slide-Rule Approximations

The slide-rule approximations for small angles exploit the instrument's logarithmic scales to simplify trigonometric computations. The sine scale (S) is positioned according to log(sin θ), where θ is typically in degrees, enabling direct reading of sin θ against the main logarithmic D scale when aligned properly. For small θ in radians, the fundamental small-angle approximation sin θ ≈ θ implies log(sin θ) ≈ θ, since sin θ / θ → 1 as θ → 0; this logarithmic identity allows users to approximate sin θ by simply reading the value of θ directly from the D scale, as the scales align closely for such angles. The technique involves setting the hairline cursor on the D scale at the position corresponding to θ (in radians, typically less than 0.3 for reasonable alignment), yielding sin θ with minimal adjustment, often treating sin θ and tan θ as interchangeable due to their near-equality for small θ. For enhanced precision, the approximation can incorporate the next term from the expansion, sin θ ≈ θ (1 - θ²/6), which translates via logarithmic expansion to log(sin θ) ≈ log θ + log(1 - θ²/6) ≈ log θ - θ²/6, adaptable on logarithmic scales for iterative refinements in analog computation. These methods were historically vital during the pre-electronic calculator period, from the 1920s to the 1970s, when slide rules served as indispensable tools in disciplines such as for rapid on-site trigonometric evaluations, including angle reductions and sight line computations. Modern analogs appear in digital tools like scientific calculators and software that compute logarithms and , enabling similar quick estimations for small angles through programmed logarithmic identities. Limitations arise primarily from the approximation's inherent error, which grows with angle size; for θ > 0.2 radians (approximately 11.5°), the relative error in sin θ ≈ θ exceeds 0.6%, compounded by the logarithmic scale's varying and , reducing readability and precision beyond the scale's typical range of 0.01 to 0.1 .

Applications

Astronomy

In astronomy, the small-angle approximation is essential for calculating to celestial objects and determining their apparent sizes, where angular measurements are typically much smaller than 1 . This approximation simplifies trigonometric relations, such as \tan \delta \theta \approx \delta \theta and \sin \alpha \approx \alpha, enabling practical computations from observed angular shifts or diameters relative to known baselines or physical scales. A primary application is the parallax method for measuring stellar distances, where the small annual shift in a star's position against background , \delta \theta, is observed over . The approximation yields \delta \theta \approx b / d, with b as the (1 ) and d the distance, directly giving d \approx 1 / p in parsecs when p = \delta \theta is in arcseconds. This holds because stellar parallaxes are minuscule, rarely exceeding 1 arcsecond for even the nearest . Modern surveys like the mission use small-angle approximations extensively for precise measurements of over 1 billion , with DR3 (2022) providing sub-millisecarcsecond accuracy for nearby objects. For angular diameters, the apparent size \alpha of an extended object like a or is approximated as \alpha \approx D / d, where D is the physical and d the , valid for small \alpha in radians. More precisely, the exact \sin(\alpha/2) = (D/2) / d simplifies to \alpha/2 \approx (D/2) / d using \sin x \approx x, which is crucial for assessing limits in observations where \alpha approaches the instrument's angular capability. In , the small-angle approximation aids analysis of perturbations to Kepler's laws, particularly for small orbital inclinations i relative to a reference plane. Here, \sin i \approx i linearizes the effects of inclination on , such as latitude variations or , allowing simplified models of nearly coplanar orbits in multi-body systems like planetary perturbations around . For example, , the nearest star to , has a of 0.768 arcseconds (as of DR3, 2022), equivalent to \theta \approx 3.72 \times 10^{-6} radians, yielding a of 1.302 parsecs via the approximation. Telescopic resolution also relies on the approximation in the Rayleigh criterion, where the minimum resolvable angle is \theta \approx 1.22 \lambda / D, with \lambda the and D the ; the small-angle form \sin \theta \approx \theta is inherent in deriving this limit for distinguishing close stellar sources. Historically, the small-angle approximation underpinned early astronomical computations, as used by in the 2nd century BCE for constructing his star catalog of over 850 stars, where precise angular separations informed rates and positional accuracies through linearized geometric relations.

Pendulum Motion

The equation of motion for a simple , consisting of a attached to a massless rod or string of length l pivoting from a fixed point under , is derived from balance as \frac{d^2 \theta}{dt^2} + \frac{g}{l} \sin \theta = 0, where \theta is the angular displacement from the vertical and g is the . For small angular displacements, the small-angle approximation \sin \theta \approx \theta (with \theta in ) simplifies this to the \frac{d^2 \theta}{dt^2} + \frac{g}{l} \theta = 0. This approximation arises from the expansion of \sin \theta = \theta - \frac{\theta^3}{6} + \cdots, neglecting higher-order terms that are negligible when |\theta| \ll 1 . The linearized equation describes simple harmonic motion, with general solution \theta(t) = \theta_0 \cos(\omega t + \phi), where the angular frequency is \omega = \sqrt{\frac{g}{l}}. Consequently, the oscillation period is T \approx 2\pi \sqrt{\frac{l}{g}}, which depends only on the pendulum length and gravity, not on the amplitude \theta_0 or mass—a feature termed isochronism that enables reliable timekeeping in devices like clocks. This independence holds under the small-angle regime, typically for initial displacements up to about 15°, where \sin \theta and \theta differ by less than 1%. However, the actual period of a pendulum increases with larger amplitudes due to the nonlinearity of \sin \theta; the fractional error in the approximate period is roughly \frac{\theta_0^2}{16} (with \theta_0 in radians), arising from the leading \theta^3 term in the expansion. For maximum angles \theta_0 < 10^\circ (approximately 0.17 radians), this error remains below 1%, making the approximation sufficiently accurate for most practical purposes. Beyond this, such as in large-amplitude swings exceeding 20°, the full nonlinear equation must be solved, often numerically or via elliptic integrals, as the motion deviates significantly from harmonicity. For instance, grandfather clocks employ pendulums with small swings (typically under 5°) to leverage the approximation for precise isochronous motion, whereas amusement park pendulum rides with swings up to 90° or more require nonlinear analysis to predict dynamics accurately. The small-angle approximation extends to lightly damped pendulums, where air resistance or friction introduces a damping term proportional to angular velocity, yielding \frac{d^2 \theta}{dt^2} + \frac{b}{m} \frac{d \theta}{dt} + \frac{g}{l} \theta = 0 (with b as the damping coefficient and m the mass). For light damping (quality factor Q \gg 1), the motion remains approximately harmonic with the same period T \approx 2\pi \sqrt{\frac{l}{g}}, but with exponentially decaying amplitude \theta(t) \approx \theta_0 e^{-\gamma t} \cos(\omega t + \phi), where \gamma = \frac{b}{2m} is small. This retains the utility of the undamped approximation in systems like precision clocks, where damping is minimized to prolong swing duration without altering the fundamental frequency significantly.

Optics

In ray optics, the paraxial approximation assumes that light rays propagate close to the optical axis, making small angles with it, typically θ ≪ 1 radian. This enables the replacement of trigonometric functions with their arguments: sin θ ≈ θ and tan θ ≈ θ, where θ is in radians. In Snell's law for refraction at an interface, n₁ sin i = n₂ sin r, the small-angle approximation simplifies to n₁ i ≈ n₂ r, facilitating linear ray tracing and first-order optical calculations. The lensmaker's formula, which relates a thin lens's focal length f to its radii of curvature R₁ and R₂ and refractive index n, derives from applying these approximations to refraction at spherical surfaces. For a ray parallel to the axis at height h from the optical axis, the small angle θ leads to h ≈ f θ at the focal point, yielding the formula (n - 1)(1/R₁ - 1/R₂) = 1/f after combining surface powers. This paraxial derivation assumes ray heights and angles remain small, ignoring higher-order terms. In Gaussian optics, which describes paraxial beam propagation, the small-angle approximation underpins the thin-lens equation 1/f = 1/o + 1/i, where o is the object distance and i the image distance. This equation models ideal imaging for collimated or focused beams with minimal divergence, treating the lens as infinitesimally thin and rays as straight lines until refraction. The approximation holds for beam waists and propagation where angles are small, enabling matrix methods for complex systems. The paraxial approximation neglects aberrations, such as , where marginal rays focus closer than paraxial rays due to differing path lengths in spherical surfaces. By restricting analysis to rays with small angles and heights, it assumes perfect focusing without such distortions, valid primarily for low systems. In practice, this limits validity to ray angles typically below 5°, beyond which errors exceed 1% in focal position. For optical instruments like microscopes and telescopes, the paraxial approximation determines focal lengths and magnifications, assuming chief rays make small angles with the axis to ensure sharp images. In a compound microscope, the objective lens forms a real intermediate image using paraxial focusing, while the eyepiece magnifies it; validity requires aperture angles under 5° to minimize off-axis errors. Similarly, telescope objectives use the approximation for distant objects, with θ < 5° ensuring accurate angular resolution without significant aberration. Fresnel diffraction, describing near-field wave patterns, employs the small-angle approximation in its integral formulation, assuming propagation angles are paraxial to simplify phase terms and derive zone plate behaviors. This contrasts with far-field Fraunhofer diffraction but shares the core angular restriction for accurate intensity predictions.

Wave Interference

In wave interference phenomena, the small-angle approximation simplifies the analysis of phase differences arising from path length variations between wave sources. For instance, in the , the path difference δ between waves from two slits separated by distance d, incident at a small angle θ to the central axis, is given by δ = d sin θ. When θ is small (typically in radians), sin θ ≈ θ, yielding δ ≈ d θ. This approximation facilitates the condition for constructive interference, where the path difference equals an integer multiple of the wavelength λ, so mλ = δ ≈ d θ for integer order m. Consequently, the angular position of the m-th bright fringe is θ_m ≈ mλ / d, and the angular spacing between adjacent fringes is Δθ ≈ λ / d. These relations hold well in the far-field regime, where the observation screen is at a large distance L compared to d, ensuring small θ values. Similar simplifications apply to diffraction gratings, which consist of multiple slits with spacing d. The grating equation for the m-th order maximum is d sin θ_m = mλ; for small θ_m, sin θ_m ≈ θ_m, reducing to θ_m ≈ mλ / d. This linear approximation aids in designing gratings for spectroscopy, where precise angular dispersion is critical for resolving wavelengths. In Young's double-slit setup, the position y of the m-th fringe on a screen at distance L is y ≈ L tan θ_m ≈ L θ_m (using tan θ ≈ θ for small θ), providing a straightforward measure of interference patterns in laboratory far-field observations. The approximation's validity diminishes for larger angles; when θ exceeds 10° (about 0.17 radians), deviations in cos θ introduce errors in intensity calculations, as the exact path difference and phase become nonlinear, potentially distorting fringe visibility. However, it remains accurate for typical laser-based setups with visible wavelengths, where θ is often below 5°. In quantum mechanics, an analogous application appears in electron diffraction, where de Broglie waves of electrons obey similar interference rules; the small-angle approximation treats the diffraction angle as θ ≈ λ_{dB} / d, with λ_{dB} = h / p (, p momentum), mirroring optical cases in experiments like Davisson-Germer.

Structural Mechanics

In structural mechanics, the small-angle approximation is fundamental to the Euler-Bernoulli beam theory, which models the deflection of slender beams under transverse loads. The theory assumes that the slope angle θ of the beam's centerline is small, allowing the approximations θ ≈ tan θ ≈ dy/dx, where y is the transverse deflection and x is the position along the beam. This simplification linearizes the geometry, neglecting higher-order terms in the strain-displacement relations. Consequently, the curvature κ of the beam is approximated as κ ≈ dθ/ds ≈ d²y/dx², where s is the arc length, enabling the governing differential equation EI d⁴y/dx⁴ = q(x) to relate the bending moment to the applied load q(x), with EI denoting flexural rigidity. These approximations hold for small deflections where rotations are typically less than 10°–15°, ensuring linear elastic behavior without significant shear deformation or axial stretching. A key application arises in stability analysis, particularly for column under axial compression. In the Euler buckling model, the small-angle approximation sin θ ≈ θ is applied to the equilibrium equation of the deflected column, leading to the differential equation d²θ/dx² + (P/EI) θ = 0, where P is the compressive load. Solving this yields the critical buckling load P_cr = π² EI / L² for a simply supported column of length L, marking the onset of instability. This formula assumes infinitesimal perturbations and small rotations, providing a conservative estimate for slender columns where the slenderness ratio L/r > π √(E/σ_y), with r as the and σ_y the yield stress. For practical examples, consider a cantilever of length L subjected to a tip load P; the small-angle approximation gives the tip deflection δ ≈ PL³ / (3EI), valid when the tip rotation θ < 15°. This result facilitates quick assessments in preliminary design, such as estimating maximum deflections under service loads. However, for larger deflections where θ exceeds this limit, the linear overpredicts δ by neglecting geometric nonlinearities like axial shortening, necessitating the von Kármán nonlinear beam . The von Kármán approach incorporates moderate rotation effects through terms like (1/2)(dy/dx)², coupling bending and stretching for more accurate predictions in post-buckling or high-load scenarios. These approximations are widely applied in engineering , such as analyzing deflections in girders under loads and structures in under aerodynamic forces, where deflections are constrained to less than 10% of the or thickness to maintain . In , they inform sizing to limit vibrations and ensure serviceability, while in , they support aeroelastic assessments under small loads. Limitations arise in extreme cases, like seismic events or high-g maneuvers, where nonlinear theories are essential to avoid underestimating stiffness. In , the small-angle approximation simplifies calculations for coordinated turns by assuming the bank angle φ is small, allowing tan φ ≈ φ (in radians). This leads to the turn radius R ≈ V² / ( φ), where V is the and is , providing pilots with a quick estimate for maneuvering without complex trigonometric computations. For typical coordinated turns, bank angles are kept below 10° to maintain accuracy of this approximation while ensuring passenger comfort and structural limits, as higher angles introduce significant errors in radius predictions. In maritime navigation, small-angle approximations apply to ship yaw angles during course corrections in calm seas, where the drift angle β is often less than 5°. Here, sin β ≈ β facilitates modeling of lateral deviations, enabling simpler trajectory predictions for systems without full nonlinear hydrodynamic simulations. Gyroscopes in systems, such as gyrocompasses, rely on small-angle approximations for analyzing due to drift. The precession angular rate ω is approximated as ω ≈ drift / time, assuming small tilt and angles where sin ε ≈ ε and cos ε ≈ 1, which uncouples the and allows for linear error modeling in heading determination. In , course deviations δ are treated similarly, with δ ≈ sin δ for small corrections, reducing the cross-track error calculation to a linear from the intended path, aiding manual plotting on charts. Inertial navigation systems (INS) integrate data to track position, using small-angle approximations for platform tilt θ to resolve accelerations into global coordinates. For small θ, the horizontal acceleration error δA ≈ ±g θ, where g is , simplifies error propagation in velocity and position updates, essential for maintaining accuracy over short periods. Gyroscopes complement this by providing updates via measurements under similar small-angle assumptions. Historically, small-angle approximations were integral to bombing sights like the Norden M-9, which compensated for small drift and trail angles in high-altitude releases to achieve precision targeting despite crosswinds. In modern applications, employing these approximations serve as backups to GPS in and piloting, providing during signal outages with bounded error growth for up to several hours.

Numerical Interpolation

In numerical methods, the small-angle approximation simplifies finite difference schemes for solving differential equations involving trigonometric functions, particularly when discretizing angular variables with small increments Δθ. For instance, the change in sine, Δ(sin θ) = sin(θ + Δθ) - sin θ, expands via Taylor series as cos θ ⋅ Δθ - (1/2) sin θ ⋅ (Δθ)^2 + O((Δθ)^3), and for small Δθ where θ is near zero, this approximates to Δθ since cos θ ≈ 1 and higher terms vanish. This linearization aids numerical solvers for oscillatory systems, reducing computational complexity while maintaining accuracy for step sizes Δθ ≪ 1 radian, with error bounded by O((Δθ)^3). For interpolation of trigonometric functions over small angular intervals, polynomial methods like Lagrange or splines exploit the near-linearity of sin θ ≈ θ and cos θ ≈ 1 - (θ^2)/2. In , fitting a through points θ_i with small |θ_i - θ_j| < ε yields a linear approximant that matches the small-angle up to second order, minimizing interpolation error for band-limited angular data. further smooths these fits, preserving monotonicity for sin θ in [0, π/6], where the approximation error remains below 0.1% for intervals up to 0.5 radians. A representative example arises in simulating two-dimensional random walks with small turning angles, where each step's direction change θ satisfies θ ≈ for modeling, simplifying path integrals from curved to straight-line approximations with error O((Δθ)^3). This is common in models, as in simulations in heterogeneous media, where θ ≈ θ linearizes the equation. In , small-angle approximations optimize loops by bypassing expensive trigonometric library calls for increments δθ < 0.01 radians, substituting δθ ≈ δθ and δθ ≈ 1 to accelerate iterative solvers by factors of 5-10 in angular coordinate updates. Such optimizations are standard in numerical libraries for applications. In modern contexts, these approximations appear in for small rotations represented by , where infinitesimal angle changes δθ yield updates q ≈ [cos(δθ/2), (δθ/2) ⋅ axis] ≈ [1, (δθ/2) ⋅ axis], enabling efficient in pipelines without full . In simulations, small fluctuations around equilibrium use sin φ ≈ φ to linearize torsional potentials, facilitating faster integration of vibrational modes in large biomolecular systems.

References

  1. [1]
    [PDF] Mathematical notes - Duke Physics
    Small angle approximations. For angle θ measured in radians, there are (Taylor) power series representing the trigonometric functions: sinθ =θ − θ3.
  2. [2]
    [PDF] Geometry and Vectors - Astronomy
    Thus if an error of less than 0.2% doesn't matter to you, you can use the small-angle approximation when θ < 0.1. If there is some purpose for which you need ...
  3. [3]
    [PDF] Math 1131 Applications: Small-Angle Approximation Fall 2019
    That sinx ≈ x for small x is called a small-angle approximation. It is illustrated numerically in the table below. The angles are in radians, so .2 = .2 radians ...Missing: physics | Show results with:physics
  4. [4]
    The Simple Pendulum - Graduate Program in Acoustics
    If the amplitude of angular displacement is small enough that the small angle approximation ( ) holds true, then the equation of motion reduces to the equation ...
  5. [5]
    Rays: Casting and Paraxial Optics
    The paraxial assumption is an example of the "small angle approximation" so dear to physics and engineering. It is a linear approximation, which describes a ...Missing: definition | Show results with:definition<|control11|><|separator|>
  6. [6]
    Small Angle Formula | Imaging the Universe - Physics and Astronomy
    The Small Angle Approximation for trigonometry states that: The equation reads, "Tangent of theta approximately equals theta". The Small Angle Formula can be ...Missing: definition | Show results with:definition
  7. [7]
    [PDF] Chapter 6 - Describing Motion in More Dimensions
    The small angle approximation will appear many times in physics. The idea is that as an angle gets smaller and smaller, the sine of the angle gets closer and ...
  8. [8]
    13.3 Visualization of Power Series Approximations - BOOKS
    This small angle approximation may be familiar to you from your study of the motion of a pendulum. Higher order terms add more higher order polynomials and ...
  9. [9]
    Small-Angle Approximation | Brilliant Math & Science Wiki
    The small-angle approximation is the term for the following estimates of the basic trigonometric functions, valid when θ ≈ 0.Missing: requirement | Show results with:requirement
  10. [10]
    15.8 Small-angle Approximation - Front Matter
    The small-angle approximation is used to approximate the values of the main trigonometric functions when the angles involved are restricted to a certain domain.
  11. [11]
    Lesson Explainer: Small-Angle Approximations | Nagwa
    In this explainer, we will learn how to approximate trigonometric functions when the angle of the function is close to zero.
  12. [12]
    Small Angle Approximations For Sin, Cos And Tan | Studywell.com
    What about larger angle values? Functions can be approximated by polynomials by taking a Taylor Expansion (what is a Taylor Expansion?).Missing: assumptions | Show results with:assumptions<|separator|>
  13. [13]
    Squeeze theorem and the limit of sin(x)/x
    The small angle approximation says that when x x x is small, sin ⁡ ( x ) ≈ x \sin(x) \approx x sin(x)≈x. This is quite useful in physics, but is imprecise.
  14. [14]
    Astronomy 102 Specials: The Observer's Triangle
    For the skinny triangle, we can calculate the length of either the arc ... This approximation is accurate to 1% for angles smaller than 9 degrees, and ...
  15. [15]
    Maclaurin Series -- from Wolfram MathWorld
    A Maclaurin series is a Taylor series expansion of a function about 0, (1) Maclaurin series are named after the Scottish mathematician Colin Maclaurin.
  16. [16]
    Taylor series | Definition, Formula, & Facts | Britannica
    Oct 24, 2025 · The series is named for the English mathematician Brook Taylor. If a = 0 the series is called a Maclaurin series, after the Scottish ...<|separator|>
  17. [17]
    Treatise of Fluxions | work by Maclaurin - Britannica
    His two-volume Treatise of Fluxions (1742), a defense of the Newtonian method, was written in reply to criticisms by Bishop George Berkeley of England.
  18. [18]
    How to justify small angle approximation for cosine
    Feb 25, 2012 · You can use the double angle formula: 1−cos2θ=2sin2θ∼2θ2. and so. cosθ∼1−θ22. or ask them to prove that. limθ→01−cosθθ2=12.Geometric Derivation of Small Angle Approximation: $\tan\theta ...A geometric proof for the "small angle approximation" for the sine ...More results from math.stackexchange.com
  19. [19]
    [PDF] Approximation and Interpolation - Numerical Analysis Lecture Notes
    The basic ideas of interpolation and least squares fitting of data can be applied to approximate complicated mathematical functions by much simpler polynomials.<|control11|><|separator|>
  20. [20]
    Sine approximation for small angles - Applied Mathematics Consulting
    Jul 27, 2010 · By contrast, saying that sin(θ) ≈ 0 for small θ is a bad approximation: the absolute error is approximately θ and the relative error is 100%.<|separator|>
  21. [21]
    6.3 Taylor and Maclaurin Series - Calculus Volume 2 | OpenStax
    Mar 30, 2016 · p n ( x ) . This theorem allows us to bound the error when using a Taylor polynomial to approximate a function value, and will be important in ...
  22. [22]
    Small Angle Approximation - Housing Innovations
    Mar 21, 2025 · The small angle approximation can be derived using the Taylor series expansion of the sine and cosine functions. The Taylor series expansion of ...
  23. [23]
    Small-Angle Approximation - (College Physics I - Fiveable
    The accuracy of the approximation decreases as the angle of oscillation increases, and using the approximation for larger angles can lead to significant errors ...
  24. [24]
    [PDF] Lecture 19: Taylor series - Berkeley Mathematics
    Apr 20, 2022 · truncating at the first term gives the small-angle approximation, but we can keep going. Since the even terms vanish, P2(x) = P1(x) = x, but ...
  25. [25]
    [PDF] Commonly Used Taylor Series
    (x − x0)n ←- the sum keeps on going and going . = P∞(x), i.e., for f to equal to its Taylor series. f(n)(x0) n! (x − x0)n ←- the sum keeps on going and going .
  26. [26]
    [PDF] Chapter 10 Series and Approximations - - Clark Science Center
    Since this limit is less than 1 for any value of x, the series converges for all x, and thus the radius of convergence of the Taylor series for ex is R = ∞, as.<|separator|>
  27. [27]
    [PDF] 11. The Trigonometric Functions
    May 9, 2002 · that sin θ/θ → 1 as θ → 0; that is, sin δ ≈ δ for small δ. But these analyses are merely numerical. A rigorous mathematical proof that sin δ ...<|control11|><|separator|>
  28. [28]
    Proving the derivatives of sin(x) and cos(x) (article) - Khan Academy
    First, we would like to find two tricky limits that are used in our proof. · Now we are ready to prove that the derivative of sin ⁡ ( x ) ‍ is cos ⁡ ( x ) ‍ .
  29. [29]
    Uncertainties and Error Propagation
    Error propagation involves calculating the uncertainty of a new quantity (z) from x and y, using a simplified statistical treatment. For addition/subtraction, ...
  30. [30]
    [PDF] The Slide Rule - UCSD CSE
    Nov 18, 2005 · sin θ ≈ tan θ, for small θ. • sin θ = b/c, tan θ = b/a. • For small θ. – a ≈ c, therefore sin θ ≈ tan θ. – Use ST scale for θ < 5.74 b a c θ ...
  31. [31]
    The Slide Rule - IEEE Pulse
    May 28, 2019 · The slide rule is an analog device, invented by William Oughtred, that uses sliding scales to perform multiplication and division.
  32. [32]
    History of the slide rule
    They were used for navigation at sea, land surveying, artillery ballistics, alcohol gauging, and carpenters calculations. Next to the usual scales for ...
  33. [33]
    Stellar Parallax - Las Cumbres Observatory
    The distance d is measured in parsecs and the parallax angle p is measured in arcseconds. This simple relationship is why many astronomers prefer to measure ...
  34. [34]
    Determining distances through parallax - University of Oregon
    But if you do know trigonometry, let me point out that all of the angles used are very small, and I use approximations valid for small angles. The apparent ...
  35. [35]
    Lecture 5: Distances of the Stars
    Sep 21, 2014 · In the upper figure, the star is about 2.5 times nearer than the star in the lower figure, and has a parallax angle which is 2.5 times larger.
  36. [36]
    Positions and Sizes of Cosmic Objects - Las Cumbres Observatory
    The Small-Angle Formula. The angular sizes of objects show how much of the sky an object appears to cover. Angular size does not, however, say anything about ...
  37. [37]
    Orbital dynamics of two circumbinary planets around misaligned ...
    Dec 2, 2021 · In evaluating the results, we calculate each planet's inclination in the small angle approximations with respect to the initial binary orbital ...
  38. [38]
    Astronomy 1144: Lecture 5 - The Ohio State University
    For example, the nearest star, Proxima Centauri, has a parallax of 0.772-arcsec (the largest parallax observed for any star). 1 arcsecond (1'') is the angle ...Missing: approximation | Show results with:approximation
  39. [39]
    Resolution | COSMOS - Centre for Astrophysics and Supercomputing
    using the small angle approximation, sin θ ≈ θ for small θ. According to the Rayleigh criterion, two point sources cannot be resolved if their separation is ...
  40. [40]
    Hipparchus - Historical Astronomy - themcclungs.net
    Hipparchus catalogs the position and brightness of over 800 stars. He was ... small 0.1 degree angle Hipparchus had found. Both angles are small enough ...
  41. [41]
  42. [42]
    [PDF] arXiv:physics/0510206v3 [physics.ed-ph] 5 Jul 2006
    Jul 5, 2006 · The small angle approximation (T ≈ T0) yields an error that is greater than 0.1% for θ0 > 7◦ and reaches 15.3% for θ0 > 90◦. The thick ...
  43. [43]
    [PDF] Numerical Solution of Linear and Nonlinear Swinging Pendulum
    Students learn what a pendulum is and how it works in the context of amusement park rides. ... large initial angular displacement cannot be solved linearly ...
  44. [44]
    Paraxial Approximation - RP Photonics
    Here, the paraxial approximation means that the angle between such rays and some reference axis of the optical system always remains small, i.e. ≪ 1 rad.
  45. [45]
    Paraxial Rays - HyperPhysics
    But for small angles, Snell's law can be approximated by n1θ1 ≈ n2θ2. The approximation tanθ ≈ θ is used in developing the expression for surface power, which ...
  46. [46]
    [PDF] Lenses and Mirrors - Duke Physics
    For paraxial rays one can show, using the law of refraction and small angle approximations, that the focal length is given by the following formula: Here n ...
  47. [47]
    [PDF] Refraction on Spherical Surface
    For small angles sn sn y y m s yn s yn b a b a. ′. −. = ′. = ′. ′. −. = R positive ... • Lensmaker's Formula. • Aberrations. – spherical – ideal lens shape is ...
  48. [48]
    Technical Note: Lens Fundamentals - Optics - Newport
    1/f = 1/s1 + 1/s2. This is the Gaussian lens equation. This equation provides the fundamental relation between the focal length of the lens and the size of the ...
  49. [49]
    2.6: Gaussian Geometrical Optics - Physics LibreTexts
    Sep 16, 2022 · All equations are also valid for a thin negative lenses and for virtual objects and images. Some examples of real and virtual object and image ...
  50. [50]
    Spherical Aberration - HyperPhysics
    For a lens with spherical aberration, the best approximation to use for the focal length is the distance at which the difference between the paraxial and ...
  51. [51]
  52. [52]
    [PDF] Section 4 Imaging and Paraxial Optics
    Paraxial Optics – A method of determining the first-order properties of an optical system by tracing rays using the slopes of the rays instead of the ray angles ...
  53. [53]
    6.6: Fresnel and Fraunhofer Approximations - Physics LibreTexts
    Sep 16, 2022 · The Fresnel and Fraunhofer approximation are two approximations of the Rayleigh-Sommerfeld integral (6.13). The approximations are based on the assumption that ...
  54. [54]
    Young's Double Slit Experiment | Physics - Lumen Learning
    The interference pattern for a double slit has an intensity that falls off with angle. ... For small angles sin θ − tan θ ≈ θ (in radians). For two adjacent ...
  55. [55]
    [PDF] Interference & Diffraction
    Analyzing Young's Double-Slit Experiment. Neil Alberding (SFU Physics) ... a sinθp = pλ,p = 1,2,3,... The small angle approximation means we can write θp ...
  56. [56]
    Young's double slit equation (video) - Khan Academy
    Apr 15, 2015 · Let's derive a formula that relates all the variables in Young's double slit experiment as we explore Young's Double Slit Equation. Uncover how light's path ...
  57. [57]
    3.3: Diffraction Gratings - Physics LibreTexts
    Oct 10, 2025 · If the slit separation is too small, then the angles between the fringes are large, resulting in very few fringes, widely separated, foiling the ...
  58. [58]
    [PDF] A Brief Introduction to Diffraction Gratings
    With the small-angle approximation, θm,i ≈ mλi/d and the angular separation between these maxima is given by ∆θm = θm,2 − θm,1 ≈ m∆λ/d. 2. With N slits there ...
  59. [59]
    Double Slit Interference - HyperPhysics
    Note: The small angle approximation was not used in the calculations above, but it is usually sufficiently accurate for laboratory calculations.
  60. [60]
    Double-Slit Interference and Diffraction Gratings – AP Physics 2 ...
    I don't understand why the small angle approximation only works when θ < 10 degrees - why that specific number? The small-angle approximation (sin θ ≈ tan θ ≈ θ ...
  61. [61]
    Lab 10
    If we observe the diffraction pattern on a screen a distance L away from the grating, then we can write dz/(mL) = λ if θ is a small angle. For electron ...<|control11|><|separator|>
  62. [62]
    [PDF] A Visualization Tool for the Vibration of Euler-Bernoulli and ...
    The signed curvature κ of the centerline is related to the angle θ as: κ = ∂θ. ∂x. (3). Thus, for small deflections of the beam, κ can be approximated by.
  63. [63]
  64. [64]
    Approximate Derivation of Critical Buckling Load - jonochshorn.com
    Feb 29, 2008 · Euler's famous equation for critical buckling load is rather cumbersome to derive, relying as it does on solving a second order differential equation.
  65. [65]
    Construction of beam elements considering von Kármán nonlinear ...
    The von Kármán nonlinear strains based formulation introduced in the paper is applicable for such class of beam problems undergoing large transverse deflection ...
  66. [66]
    [PDF] A STUDY OF LARGE DEFLECTION OF BEAMS AND PLATES
    This model, Euler-Bernoulli beam theory, does not consider the effect of rotatory iner- tia and the correction for transverse shear. The nonlinear beam theory ...
  67. [67]
  68. [68]
    Page Not Found
    Insufficient relevant content. The provided URL (https://www.faa.gov/regulations_policies/handbooks_manuals/aviation/airplane_flying_handbook/media/afh_ch05.pdf) results in a "Page Not Found" error, and the accessible content does not contain information on bank angles in coordinated turns for aircraft or typical values for small angles. The page only includes general website navigation and security information.
  69. [69]
    Ship Motion Control Methods in Confined and Curved Waterways ...
    For restricted environment, the drift angle β is typically small (|β| < 5°), justifying the small-angle approximation (cosβ ≈ 1, sinβ ≈ β). The derivative ...Ship Motion Control Methods... · 3. Ship Motion Control... · 3.4. Optimal Pid Control...
  70. [70]
    [PDF] matical Models for Gyrocompass Behaviour: Error Modelling and ...
    very close to 90°) the period of precession of the gyrocompass becomes infinite. ... for small angle approximation, cos £ ~ 1 and sin £ = £. (I-7) thus. aT. (A+B ...
  71. [71]
    [PDF] Fundamentals of GNSS-Aided Inertial Navigation - IntechOpen
    Jan 18, 2012 · using the small angle approximation of (20), where sin ϕ ≈ ϕ, sin θ ≈ θ and sin ψ ≈ ψ, which is. Rn b ≈ ⎛. ⎝. 1 −ψ θ ψ 1 −ϕ. −θ ϕ. 1.
  72. [72]
    [PDF] CHAPTER 21 INTRODUCTION TO INERTIAL NAVIGATION
    The small angle approximation is used to simplify the error term. The positive or negative sign is determined by the assumed direction of motion over the ...
  73. [73]
    Norden M-9 Bombsight - Air Force Museum
    The Norden bombsight was crucial to the success of the US Army Air Forces' daylight bombing campaign during World War II.Missing: angle approximation
  74. [74]
    [PDF] lecture 7: Trigonometric Interpolation - Virginia Tech
    The coefficient 7` that premultiplies ei`x in the trigonometric interpolating polynomial is actually an approximation of the Fourier coefficient c`. Let us go ...Missing: angles | Show results with:angles
  75. [75]
    Interpolating sines and cosines - Applied Mathematics Consulting
    Oct 3, 2021 · You can beat linear interpolation for filling in gaps in a table of sines and cosines by using trig identities for adding angles.
  76. [76]
    An Improved Small-Angle Approximation for Forward Scattering and ...
    Jun 1, 2017 · For the diffuse component, the source and the forward-scattering matrix are replaced by Dirac delta function, which can cause noticeable errors ...
  77. [77]
    [PDF] arXiv:astro-ph/9707095v1 8 Jul 1997
    Angular skewness s3 for galaxy catalogue selection function from Monte-Carlo integration (symbols) compared with small angle approximation PT predictions (lines) ...
  78. [78]
    Fast and Accurate Approximation Methods for Trigonometric ... - MDPI
    Jul 22, 2022 · In this study, we designed an algorithm for the approximate computation of trigonometric and inverse trigonometric functions, which are nonlinear elementary ...
  79. [79]
    [PDF] Representing and Estimating Rotations in R3
    the small angle approximation. More on Rotations.