The Gauss circle problem is a classical question in analytic number theory that seeks to determine the number N(r) of integer lattice points (m, n) \in \mathbb{Z}^2 lying inside or on the boundary of a circle of radius r > 0 centered at the origin, satisfying m^2 + n^2 \leq r^2. This count N(r) is asymptotically approximated by the area of the circle, \pi r^2, with the discrepancy given by the error term E(r) = N(r) - \pi r^2; the problem focuses on obtaining the sharpest possible bounds for |E(r)| as r \to \infty.[1]Named after the mathematician Carl Friedrich Gauss, who first considered it around 1800, the problem has deep connections to the distribution of lattice points, exponential sums, and modular forms, and it generalizes to higher dimensions and other shapes. Gauss established an initial elementary bound |E(r)| \leq 2\sqrt{2\pi} r, reflecting the influence of the circle's circumference.[2] Over the subsequent two centuries, significant progress has been made through the works of mathematicians including Sierpiński, Hardy, Landau, van der Corput, Vinogradov, and Huxley, who refined the upper bounds using advanced techniques from Fourier analysis and estimates on oscillatory integrals.[3]Despite these advances, the problem remains unsolved, with a central conjecture asserting that E(r) = O(r^{1/2 + \epsilon}) for every \epsilon > 0, implying the error is essentially of order the square root of the radius up to logarithmic factors. In 1915, Hardy and Landau proved a matching lower bound in the sense that E(r) \neq o(r^{1/2} (\log r)^{1/4}), showing that the conjectured order is essentially optimal. The best known upper bound to date is |E(r)| = O(r^{131/208 + \epsilon}) (where $131/208 \approx 0.6298), achieved by Huxley in 2003 via improvements in bounding exponential sums over curves.[3] This exponent has seen only marginal refinements since, such as a slight improvement to $517/824 + \epsilon \approx 0.627 + \epsilon proposed by Bourgain and Watt in 2017 (though the paper was later withdrawn), underscoring the challenge in breaking below the 0.63 barrier without new methods. The problem's resolution would have implications for related questions, including the Dirichlet divisor problem and lattice point discrepancies in other domains.[4]
Historical Development
Origins and Gauss's Contribution
The Gauss circle problem traces its origins to the early 19th century, when Carl Friedrich Gauss began studying the distribution of integer lattice points within circles as part of his investigations into quadratic forms and the representation of integers as sums of two squares. These efforts were motivated by arithmetic questions, particularly the asymptotic behavior of the function r_2(n), which counts the number of ways to express n as a sum of two integer squares, leading Gauss to consider lattice point counts in the plane.Gauss's foundational work on the problem remained unpublished during his lifetime, preserved instead in his private manuscripts and notes from approximately 1800 to 1810. In these documents, he analyzed the number N(r) of lattice points (m, n) \in \mathbb{Z}^2 satisfying m^2 + n^2 \leq r^2, deriving the asymptotic relation N(r) = \pi r^2 + E(r), where E(r) denotes the discrepancy or error term. Through geometric arguments centered on the circle's perimeter, Gauss obtained the initial bound |E(r)| \leq 2 \sqrt{2} \pi r, which estimates the contribution of points near the boundary by considering annuli of width \sqrt{2} around the circle.[2]This bound provided an early insight into the problem's scale, linking it to Diophantine approximation and the even distribution of lattice points modulo 1. Gauss's approach reflected his broader interests in number theory, with related ideas appearing in his mathematical diary as early as 1799, where he explored arithmetic-geometric means and their analytic implications.[5] For small radii, such as r = 1, Gauss performed explicit counts, finding N(1) = 5 lattice points: the origin and the four adjacent axis points.[6]
Key Early Results
Building upon Gauss's initial estimate of |E(r)| \leq 2 \sqrt{2} \pi r, early 20th-century advancements shifted focus toward sharper upper bounds using emerging analytic techniques. In 1903, Georgy Voronoi established the classical bound |E(r)| = O(r^{2/3}) using methods from the theory of quadratic forms.[2] This was independently confirmed in 1906 by Wacław Sierpiński, who improved the constant to |E(r)| < 2.58 r^{2/3}, marking a significant step beyond the linear barrier by leveraging asymptotic analysis of lattice point distributions. This bound, derived through detailed examination of the error in the circle's area approximation, highlighted the potential for exponents between $1/2 and 1 in the error term.[7]Further progress came in 1923 with J. G. van der Corput's application of exponential sums, yielding the classical bound |E(r)| = O(r^{2/3}). Van der Corput's method, introduced in his work on number-theoretic estimates, exploited oscillatory integrals to control the discrepancy between the lattice count and the geometric area, establishing $2/3 as a foundational exponent for the problem. This approach not only refined previous constants but also demonstrated the power of Fourier-analytic tools in bounding lattice point errors.[8]In 1915, G. H. Hardy proved that |E(r)| \neq O(r^{1/2}), showing the existence of infinitely many r such that |E(r)| \gg r^{1/2}. Concurrently, Edmund Landau obtained the \Omega-result E(r) \neq o(r^{1/2} (\log r)^{1/4}), demonstrating that the error term achieves at least this magnitude infinitely often and underscoring the sharpness of the conjectured O(r^{1/2 + \epsilon}) bound. These lower-bound results, rooted in analytic continuations and residue calculus, revealed inherent limitations in approximating the circle's lattice points.[9]These developments occurred amid broader advances in analytic number theory during the 1910s and 1920s, linking the Gauss circle problem to the Riemann zeta function and Dirichlet's divisor problem through shared techniques like contour integration and mean-value estimates. The circle problem's error term parallels the divisor function's discrepancy, with both benefiting from zeta-function approximations to probe average orders and oscillations.[9]
Problem Formulation
Definition and Notation
The integer lattice in the Euclidean plane consists of all points (m, n) where m, n \in \mathbb{Z}. These points represent the vertices of unit squares tiling the plane. The Euclidean norm of such a point, \|(m, n)\| = \sqrt{m^2 + n^2}, measures its distance from the origin (0,0).The Gauss circle problem concerns the number of lattice points lying inside or on the boundary of the disk of radius r > 0 centered at the origin.[2] This number is denoted N(r) and formally defined asN(r) = \# \bigl\{ (m,n) \in \mathbb{Z}^2 : m^2 + n^2 \leq r^2 \bigr\},where \#\{\cdot\} denotes cardinality.[2] Equivalently, N(r) counts the lattice points (m,n) satisfying \|(m,n)\| \leq r.[10]The area of the disk is \pi r^2, providing a natural approximation for N(r) under the assumption of uniform distribution.[11] Thus, N(r) \approx \pi r^2, and the difference E(r) = N(r) - \pi r^2 quantifies the discrepancy, often called the error term.[2]N(r) is a non-decreasing step function in r: it remains constant for intervals of r and jumps upward precisely when r^2 passes an integer that is a sum of two squares, with the jump size equal to the number of distinct lattice points at that distance (accounting for symmetries).[2] For instance, when r=2 (r^2=4), N(2)=13, compared to the area \pi \cdot 4 \approx 12.566.[2] The problem was motivated by early 19th-century investigations into lattice point distribution by Carl Friedrich Gauss.[2]
The Error Term
The error term in the Gauss circle problem is defined as E(r) = N(r) - \pi r^2, where N(r) denotes the number of integer lattice points (m,n) \in \mathbb{Z}^2 satisfying m^2 + n^2 \leq r^2. This term quantifies the discrepancy between the discrete lattice count and the continuous area of the disk. It possesses key properties: |E(r)| grows sublinearly, satisfying E(r) = o(r) as r \to \infty. The conjectured sharp upper bound is O(r^{1/2 + \epsilon}) for every \epsilon > 0, and Hardy and Landau (1915) showed that |E(r)| = \Omega(r^{1/2} (\log r)^{1/4}) infinitely often, indicating that the exponent $1/2$ is optimal up to logarithmic factors.[2][12]The error term E(r) displays pronounced oscillatory behavior, changing sign infinitely often with regularly spaced transitions in its normalized form E(r)/\sqrt{r}. This oscillation arises from the irregular alignment of lattice points near the circle's boundary, leading to alternating over- and under-counts relative to the area \pi r^2. Mean square estimates further characterize its average behavior, with \int_0^r E(t)^2 \, dt \sim c r^{3/2} for some constant c > 0, implying a root mean square order of O(r^{1/4}), as shown by Hardy and Littlewood (1922).[13][14]Numerical examples illustrate these fluctuations on small scales. For r=10, E(10) \approx 2.84, indicating a modest overcount. For r=100, E(100) \approx 1.07, though the error exhibits fluctuations with magnitudes growing sublinearly, underscoring the growing variability despite the sublinear overall growth.[11]In discrepancy theory, the normalized error E(r)/r measures the lattice's approximation quality by the circle, effectively gauging the average deviation per unit length along the boundary arc of length $2\pi r. This perspective connects the Gauss circle problem to uniform distribution problems, where large |E(r)/r| signals poor alignment between the lattice and the curved domain.[15]
Asymptotic Analysis
Classical Bounds
The classical bounds on the error term E(r) in the Gauss circle problem, defined as the difference between the number of lattice points inside the circle of radius r centered at the origin and the area \pi r^2, were established using elementary and early analytic methods. Around 1800 (unpublished during his lifetime), Carl Friedrich Gauss provided the first non-trivial upper bound via a perimeter argument, estimating the discrepancy by the number of lattice points near the circle's boundary. Specifically, Gauss showed that |E(r)| \leq 2\sqrt{2} \pi r \approx 8.9 r.[2]Significant refinements to the upper bounds emerged in the early 20th century. In 1906, Wacław Sierpiński improved the estimate to |E(r)| < c r^{2/3} for some constant c > 0, marking the first sublinear exponent beyond Gauss's O(r). This r^{2/3} bound, equivalent to an exponent \theta = 2/3 \approx 0.6667, was independently confirmed by J. G. van der Corput in 1923 using exponential sum techniques. Van der Corput further advanced the result in 1922 by achieving |E(r)| = O(r^{33/50}), where $33/50 = 0.66, and in 1928 refined it to O(r^{27/41}) with $27/41 \approx 0.6585. These early 20th-century improvements, including subsequent minor adjustments in the 1920s, established \theta < 2/3 as the classical frontier before more sophisticated analytic tools.[14][10]On the lower bound side, Edmund Landau's 1915 analysis demonstrated that the error term cannot be smaller than the conjectured order in a strong sense. He proved that \limsup_{r \to \infty} |E(r)| / (r^{1/2} (\log r)^{1/4}) > 0, implying the existence of infinitely many r where |E(r)| is non-trivially large, at least on the scale of r^{1/2} times a logarithmic factor. This result, obtained through properties of L-functions and zeta zeros, underscored the problem's depth and complemented the upper bounds by showing \theta \geq 1/2.[16]
The Conjecture
The central unsolved conjecture in the Gauss circle problem, known as Hardy's conjecture, posits that the error term satisfies |E(r)| = O(r^{1/2 + \epsilon}) for every \epsilon > 0. This bound was first conjectured by G. H. Hardy in 1915, motivated by the structural analogy between the circle problem and the Dirichlet divisor problem, where the corresponding error is expected to be O(x^{1/4 + \epsilon}).[9]The conjecture arises from heuristic considerations, including probabilistic models that treat lattice points as randomly distributed, predicting an error on the order of the square root of the circle's circumference, approximately \sqrt{2\pi r}, and principles of equidistribution that suggest the fractional parts of lattice point coordinates behave uniformly. Achieving this bound would demonstrate that the area \pi r^2 approximates the number of lattice points exceptionally well, with the discrepancy negligible relative to r^{1/2 + \epsilon}.[17]Proving the conjecture would resolve key open questions in analytic number theory, particularly by establishing the optimal exponent for the Dirichlet divisor problem due to the deep interconnections between the two via representation functions and Dirichlet series.[9]While the pointwiseconjecture remains open, it holds in an average sense, as evidenced by the mean square estimate \int_0^R E(r)^2 \, dr \sim c R^2 \log R for some constant c > 0, which aligns with the expected magnitude under the conjecture. Additionally, E. Landau established in 1915 that the exponent $1/2 is optimal, showing |E(r)| = \Omega(r^{1/2}).
Modern Improvements
Significant progress on bounding the error term E(r) in the Gauss circle problem has been made since the mid-20th century through advanced analytic number theory techniques, particularly involving exponential sums and sieve methods. In 1988, Iwaniec and Mozzochi established the bound |E(r)| = O(r^{7/11 + \epsilon}) for any \epsilon > 0, where $7/11 \approx 0.6364, utilizing refinements of the circle method combined with estimates for Farey fractions and van der Corput's treatment of exponential sums. This improved upon earlier exponents around 0.648 from the 1960s and 1970s, such as Kolesnik's 1982 result of O(r^{35/54 + \epsilon}) with $35/54 \approx 0.6481.Further advancements came in the 1990s and early 2000s with Huxley's work on exponential sums and lattice points. In 1993, Huxley achieved |E(r)| = O(r^{46/73 + \epsilon}) where $46/73 \approx 0.6301, employing mean value theorems for Dirichlet polynomials. His 2003 refinement to |E(r)| = O(r^{131/208 + \epsilon}) with $131/208 \approx 0.6298 incorporated the Bombieri-Iwaniec method for handling major arcs and a square-free sieve to control minor arc contributions, marking the current best unconditional bound.Modern techniques underpinning these bounds rely heavily on Fourier analysis via the Poisson summation formula to express E(r) in terms of Gaussian sums, the Hardy-Littlewood circlemethod to dissect the problem into major and minor arcs, and Weyl differencing to estimate higher-degree exponential sums. Additionally, representations from GL(2) automorphic forms have been explored to link the error term to spectral properties on modular surfaces. Recent efforts connect the problem to quantum ergodicity, where equidistribution of eigenfunctions on arithmetic surfaces informs bounds on lattice point discrepancies.As of 2025, ongoing research continues to push the exponent below 0.63. A talk at the AMS Spring Central Sectional Meeting highlighted improvements via estimates on L^p norms of exponential sums, leveraging decoupling theory, though specific exponents remain conjectural and unconfirmed in print. These developments suggest potential sub-0.62 bounds, maintaining the conjecture |E(r)| = O(r^{1/2 + \epsilon}) as the ultimate target.[18]
Exact Representations
Summation Formulas
One exact representation for the number of lattice points N(r) inside or on the circle of radius r centered at the origin is given by the summation formulaN(r) = \sum_{k=0}^{\lfloor r^2 \rfloor} r_2(k),where r_2(k) denotes the number of integer solutions (x, y) to x^2 + y^2 = k.[19]Here, r_2(k) counts all representations, including orders and signs, so r_2(0) = 1 for the origin and r_2(k) = 0 if k cannot be expressed as a sum of two squares.[19] This function satisfies the explicit formular_2(k) = 4 \bigl( d_1(k) - d_3(k) \bigr),where d_i(k) is the number of positive divisors of k congruent to i modulo 4.[20] Equivalently,r_2(k) = 4 \sum_{d \mid k} \chi(d),with \chi the non-principal Dirichlet character modulo 4, defined by \chi(d) = 0 if d is even, \chi(d) = 1 if d \equiv 1 \pmod{4}, and \chi(d) = -1 if d \equiv 3 \pmod{4}.[19]This summation arises directly from grouping lattice points by their squared distance k = x^2 + y^2 from the origin, with each k \leq r^2 contributing exactly r_2(k) points.[19] For instance, when r = \sqrt{2}, \lfloor r^2 \rfloor = 2, and N(\sqrt{2}) = r_2(0) + r_2(1) + r_2(2) = 1 + 4 + 4 = 9, accounting for all nine lattice points with coordinates in \{-1, 0, 1\}.The partial sums of r_2(k) connect to the asymptotic behavior of N(r), as Jacobi's theorem establishes that\sum_{k=1}^m r_2(k) = \pi m + O(\sqrt{m})for large m, implying \sum_{k=0}^{\lfloor r^2 \rfloor} r_2(k) \sim \pi r^2.[19]
Integral and Series Expressions
An elementary exact representation for N(r) isN(r) = 1 + 4 \sum_{k=1}^{\lfloor r \rfloor} \left\lfloor \sqrt{r^2 - k^2} \right\rfloor.This formula arises from counting points in the first quadrant and multiplying by 4, plus the origin, by fixing the x-coordinate and finding the maximum y.[2]One exact representation for the number of lattice points N(r) inside or on the circle of radius r centered at the origin is given by the alternating floor function seriesN(r) = 1 + 4 \sum_{i=1}^{\lfloor r^2 \rfloor} (-1)^{i-1} \left\lfloor \frac{r^2}{2i-1} \right\rfloor.This alternating sum converges rapidly for computational purposes, as terms diminish after $2i-1 > r^2. A refined non-alternating form, which separates contributions based on quadratic residues modulo 4, isN(r) = 1 + 4 \sum_{i=0}^\infty \left( \left\lfloor \frac{r^2}{4i+1} \right\rfloor - \left\lfloor \frac{r^2}{4i+3} \right\rfloor \right).These expressions allow efficient exact computation of N(r) for moderately large r by truncating the sum at terms where the floor functions vanish.[2]Poisson summation formula applied to the characteristic function of the disk yields exact integral and series representations involving special functions. Specifically, relating N(r) to the cumulative sum of the representation function r_2(n) (the number of ways to write n as a sum of two squares), Hardy's identity provides\sum_{n=0}^{\lfloor x \rfloor} r_2(n) = \pi x + \sum_{n=1}^\infty \frac{r_2(n)}{\sqrt{n}} \sqrt{x} \, J_1 \left( 2\pi \sqrt{n x} \right),where J_1 is the Bessel function of the first kind of order 1, and x = r^2. Adjusting for the boundary term givesN(r) - \frac{1}{2} r_2(\lfloor r^2 \rfloor + 1) = \pi r^2 + r \sum_{n=1}^\infty \frac{r_2(n)}{\sqrt{n}} J_1 \left( 2\pi r \sqrt{n} \right).The oscillatory nature of the Bessel functions J_1(z) \sim \sqrt{2/(\pi z)} \cos(z - 3\pi/4) for large z ensures rapid decay of the series tail, enabling precise numerical evaluation for large r via truncation after a few terms.[21]These analytic expressions, derived from summation techniques, contrast with discrete summation formulas by incorporating continuous integrals and special functions, facilitating both theoretical analysis and high-precision computations.
Extensions and Variants
Higher-Dimensional Analogues
The higher-dimensional analogues of the Gauss circle problem extend the counting of lattice points from two-dimensional disks to balls in \mathbb{R}^d for d > 2. Define N_d(r) = \# \{ \mathbf{x} \in \mathbb{Z}^d : \|\mathbf{x}\|_2 \leq r \} as the number of integer lattice points inside or on the boundary of the Euclideanball of radius r centered at the origin. The volume of this ball is V_d r^d, where V_d = \frac{\pi^{d/2}}{\Gamma(d/2 + 1)} is the volume of the unitball in \mathbb{R}^d. The error term is then E_d(r) = N_d(r) - V_d r^d, which measures the discrepancy between the lattice point count and the continuous volume approximation.[13][22]In three dimensions, the problem—often called the Gauss sphere problem—remains particularly challenging. It is conjectured that |E_3(r)| = O(r^{1 + [\epsilon](/page/Epsilon)}) for any \epsilon > 0, reflecting an expected near-optimal error relative to the surface area scale of O(r^2). The current best known bound is |E_3(r)| = O(r^{4/3 + [\epsilon](/page/Epsilon)}), achieved through advanced analytic techniques but still falling short of the conjecture. For d \geq 4, the exponents in the error bounds improve progressively; for instance, the error term satisfies |E_d(r)| = O(r^{d-2 + [\epsilon](/page/Epsilon)}) in general, with the relative discrepancy shrinking as d increases due to enhanced cancellation effects and better volume approximations in higher dimensions.[13][13][13]Key techniques for these analogues generalize the Hardy-Littlewood circle method to higher dimensions, though its efficacy diminishes in d=3 owing to insufficient major arc contributions. For d \geq 4, stronger results arise from connections to Waring's problem, which facilitates estimates on representation numbers of integers as sums of d squares, aiding in the decomposition of the error term. In the case d=2, the formulation reduces precisely to the original Gauss circle problem. For d=4, exact formulas for N_4(r) can be derived using theta functions, such as the fourth power of the Jacobi theta series \theta_3(z)^4, which encodes the lattice point counts through modular form properties.[13][13][13]
Non-Circular Domains
The elliptic analogue of the Gauss circle problem concerns the number of integer lattice points (m, n) satisfying (m/a)^2 + (n/b)^2 \leq [1](/page/1) for positive real numbers a and b. The expected main term is the area \pi a b, and the conjectured error term is O((a b)^{1/2 + \epsilon}) for every \epsilon > [0](/page/0), analogous to the circle case where a = b = r. This conjecture appears in studies of lattice point distribution in thin elliptic annuli, where Bleher and Cheng proposed bounds on the variance of the error function for generic ellipses with irrational eccentricity. For specific rational cases, such as when the ellipse aligns with lattice symmetries, the error can exhibit periodic behavior tied to modular forms.[23][24]A hyperbolic analogue arises in the context of indefinite binary quadratic forms, counting lattice points in regions like |x^2 - y^2| \leq r, which form hyperbolic strips rather than bounded domains. This problem links to the geometry of numbers for indefinite forms, where the number of representations grows logarithmically due to the unbounded nature, and the discrepancy relates to the class number of quadratic fields. Unlike the definite case, spectral methods from automorphic forms provide asymptotics, with error terms involving the divisor function and bounds like O(r^{1/2 + \epsilon}) under the Lindelöf hypothesis. Such counts appear in applications to geodesic flows on hyperbolic surfaces and visibility problems in non-Euclidean lattices.[25][26]For polygonal domains, particularly rational polytopes in the plane, Ehrhart theory establishes that the number of lattice points in the t-dilation tP is given exactly by a quasi-polynomial L_P(t) of degree 2, where the coefficients are periodic functions with period dividing the denominator of the polytope's vertices. This quasi-polynomial captures the arithmetic structure, with the leading term being the area of P and lower terms involving boundary and vertex data. For example, in rational triangles, the Ehrhart quasi-polynomial simplifies due to the low dimension, allowing explicit computation via triangulation.[27][28]Van der Corput's exponential sum methods extend to general convex domains with smooth boundaries, yielding discrepancy bounds of O(r^{2/3}) for the error in approximating the lattice point count by the area, where r scales the domain size; this holds for strongly convex curves with positive Gaussian curvature, improving on earlier O(r^{1/2 + \epsilon}) estimates. For polygonal boundaries lacking smoothness, the error can be larger, but exact relations like Pick's theorem provide precise counts for simple cases: for a latticepolygon with I interior points and B boundary points, the area is A = I + B/2 - 1, allowing inversion to determine I exactly from area and boundary data in triangles. This theorem, originally proved via Euler characteristic, underpins discrepancy analysis for non-smooth domains like triangles, where the error vanishes exactly for primitive cases.[29][30]
Primitive Lattice Points
The primitive lattice points in the Gauss circle problem are the points (m, n) \in \mathbb{Z}^2 satisfying \gcd(m, n) = 1 and m^2 + n^2 \leq r^2, excluding the origin (0,0) since \gcd(0,0) is undefined. The quantity V(r) denotes the number of such points. Asymptotically, V(r) = \frac{6 r^2}{\pi} + E'(r), where the leading term arises from the area \pi r^2 of the disk multiplied by the natural density \frac{6}{\pi^2} of primitive lattice points in \mathbb{Z}^2, given by the reciprocal of the Riemann zeta function at s=2 via \zeta(2)^{-1} = \frac{6}{\pi^2}.An exact expression for V(r) follows from Möbius inversion applied to the indicator function of coprimality: since \sum_{d \mid \gcd(m,n)} \mu(d) = 1 if \gcd(m,n)=1 and $0$ otherwise, it follows thatV(r) = \sum_{\substack{m,n \in \mathbb{Z} \\ m^2 + n^2 \leq r^2 \\ (m,n) \neq (0,0)}} \sum_{d \mid \gcd(m,n)} \mu(d) = \sum_{d=1}^{\lfloor r \rfloor} \mu(d) \left( N\left( \frac{r}{d} \right) - 1 \right),where N(s) is the total number of lattice points (including the origin) inside or on the circle of radius s, and the sum is effectively finite since \mu(d) = 0 for d > r and N(s) = 1 for s < 1. This formula isolates the primitive count by subtracting the origin's contribution in each scaled disk.The error term E'(r) = V(r) - \frac{6 r^2}{\pi} in the primitive circle problem satisfies the conjectured bound |E'(r)| = O(r^{1/2 + \varepsilon}) for any \varepsilon > 0, analogous to the non-primitive case, reflecting the belief that the discrepancy arises primarily from the boundary fluctuations of the disk. Assuming the Riemann hypothesis, this is strengthened, but the best unconditional bound is |E'(r)| = O(r^{0.31448 + \varepsilon}) as of 2023, obtained via estimates on exponential sums over lattice points near curves, mirroring the exponent for the total count N(r). This improves upon the prior bound of O(r^{131/416 + \varepsilon}) \approx O(r^{0.3149 + \varepsilon}) from Huxley's work and earlier estimates such as O(r^{37/118 + \varepsilon}). Recent unconditional improvements for the primitive error parallel those for N(r), achieving exponents below $1/3 in some mean-square settings, though the pointwise bound remains governed by the 2023 result.[31]