Next Article in Journal
Charge Density Waves in Solids—From First Concepts to Modern Insights
Previous Article in Journal
Efficient Outsourced Decryption System with Attribute-Based Encryption for Blockchain-Based Digital Asset Transactions
Previous Article in Special Issue
Gaussian Chebyshev Polynomials and Their Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

From Complex to Quaternions: Proof of the Riemann Hypothesis and Applications to Bose–Einstein Condensates

by
Jau Tang
Institute of Technological Sciences, Wuhan University, Wuhan 430074, China
Symmetry 2025, 17(7), 1134; https://doi.org/10.3390/sym17071134
Submission received: 19 June 2025 / Revised: 4 July 2025 / Accepted: 9 July 2025 / Published: 15 July 2025

Abstract

We present novel proofs of the Riemann hypothesis by extending the standard complex Riemann zeta function into a quaternionic algebraic framework. Utilizing λ-regularization, we construct a symmetrized form that ensures analytic continuation and restores critical-line reflection symmetry, a key structural property of the Riemann ξ(s) function. This formulation reveals that all nontrivial zeros of the zeta function must lie along the critical line Re(s) = 1/2, offering a constructive and algebraic resolution to this fundamental conjecture. Our method is built on convexity and symmetrical principles that generalize naturally to higher-dimensional hypercomplex spaces. We also explore the broader implications of this framework in quantum statistical physics. In particular, the λ-regularized quaternionic zeta function governs thermodynamic properties and phase transitions in Bose–Einstein condensates. This quaternionic extension of the zeta function encodes oscillatory behavior and introduces critical hypersurfaces that serve as higher-dimensional analogues of the classical critical line. By linking the spectral features of the zeta function to measurable physical phenomena, our work uncovers a profound connection between analytic number theory, hypercomplex geometry, and quantum field theory, suggesting a unified structure underlying prime distributions and quantum coherence.

1. Introduction

Riemann’s hypothesis (RH), first formulated in 1859 by German mathematician B. Riemann, is one of the most profound and long-standing unsolved problems in mathematics [1,2,3,4,5,6,7]. He postulated that the nontrivial zeros of the Riemann zeta function ζ(s) must lie along the critical line in the complex plane s = 1 2 + i y [1]. This zeta function is deeply connected to the distribution of prime numbers, forming the foundation of modern analytic number theory. This hypothesis is one among the list of 23 unsolved problems presented by D. Hilbert in 1900 at the International Congress of Mathematicians [3,4]. Despite numerous partial results obtained by notable mathematicians, such as Hardy [5], Selberg [6], Speiser [7], and many others, and an astronomical number of zeros computationally identified with a zero having an imaginary part as large as 8.1 × 1034 [8], the RH remains unsolved [9] since Riemann proposed it more than a century ago. However, the research on the mathematical properties of the zeta function and its applications in physics has remained an active subject [10,11,12,13,14,15,16,17].
The proof of the RH has far-reaching implications across number theory [18], statistical mechanics [19,20], quantum theory [21,22], and random matrix theory for chaos [23]. We present here an elegant and rigorous proof of RH. Our approach is based on the analysis of the reflection symmetry between | Γ s / 2 ζ ( s ) / π s / 2 | 2 and Γ ( 1 s / 2 ) ζ 1 s / π 1 s / 2 | 2 to establish the validity of Riemann’s conjecture.
This work takes a fundamentally new approach by extending the classical complex zeta function into a quaternionic algebraic structure. We introduce a λ-regularized and symmetrized zeta function defined over quaternionic variables. This formulation not only preserves the essential features of the original zeta function but also restores and enforces symmetry across the critical line. The non-associative geometry of hypercomplex numbers, particularly quaternions, plays a pivotal role in the emergence of this structure.
The main result of this paper is a constructive and algebraically rigorous proof of the Riemann hypothesis using this quaternionic framework. Beyond the realm of pure mathematics, we also demonstrate that this formulation has significant implications in quantum statistics. Specifically, we explore its application to Bose–Einstein condensates (BECs) [24,25], showing that the λ-regularized quaternionic zeta function governs thermodynamic properties and phase transitions in these systems. This dual-purpose approach—resolving the RH while revealing physical consequences—suggests a deep and previously unrecognized unity between prime number theory, hypercomplex analysis, and quantum statistical mechanics.

2. Two Proof Approaches to Riemann’s Hypothesis

To prove RH, we first review basics in Section 2.1 and the first approach based on Riemann’s xi function in Section 2.2. Then, in Section 2.2, we present a second approach based on the λ-regularized zeta function. Such a method allows us to generalize the standard Riemann zeta function on the complex plane to higher-dimensional hypercomplex structures, such as quaternions, octonions, and sedenions [26,27,28,29]. In this work, we shall only consider the applications of complex and quaternionic zeta functions to quantum statistics such as the Bose–Einstein condensates and phase transitions.

2.1. Basics of the Riemann Zeta Function, Xi Function, and Dirichlet Series

In this section, we first outline some basics of Euler’s zeta function, Dirichlet series, and Riemann’s zeta and xi functions. We shall present our regularized composite functions to analyze the symmetry and convexity to prove Riemann’s hypothesis and to expand the complex domain to 4D quaternions. Before the work of Riemann [1,2,3], the zeta function in Euler’s era is defined as
ξ ( x ) = 1 Γ x 0 d z z s 1 e z 1 = n = 1 1 n x ,
where x is real and the Dirichlet series form converges only for Re(s) > 1. Euler demonstrated xn interesting relation between the zeta function to a product of terms involving all prime numbers, as shown by
ζ ( s ) = p : p r i m e       1 1 p s
Riemann extended the zeta function to the complex plane via analytic continuation and formulated
ζ ( s ) = 1 Γ s 0 d x x s 1 e x 1
where s is complex and ξ ( s ) is analytic except x = 1. He further showed
ζ ( s ) = 2 s π s 1 s i n π s / 2 Γ 1 s ζ 1 s
According to RH [4,5,6], the zeros of the zeta function occur only along the critical line with x = 1/2. Because it is well known that the zeros of the Riemann zeta function occur along the critical strip with x between 0 and 1 [5], to prove RH, one only needs to analyze the location of the minimum for in the critical strip, which happens to be at the zeros if the zeta function must lie along the critical line.
However, the Dirichlet series representation of the zeta function diverges for Re(s) ≤ 1. To extend its convergence domain to the critical strip, we include an exponential damping term to construct the regularized zeta function, defined as
ζ λ ( s ) = n = 1   e n λ / n s
This λ-regularized function converges on the whole complex plane and has structural similarity to partition functions in quantum statistics. One can apply such a regularization procedure to circumvent the divergence at x = 1 for Riemann’s ζ ( s ) in Equation (3) and obtain
ζ λ ( s ) = 1 Γ s 0 d x x s 1 e x + λ 1 = n = 1 e n λ n s
where the regularization parameter can be related to the physical quantity called fugacity. The kernel inside the integral represents the Bose–Einstein statistics; therefore, this λ-regularized zeta function is intrinsically related to Bose–Einstein condensates, which we will address in later sections.
We would like to point out that if we consider the l-regularized eta function, instead of the zeta function, we can obtain Fermi–Dirac statistics. Here we define
η λ ( s ) = 1 Γ s 0 d x x s 1 e x + λ + 1 = n = 1 1 n + 1 e n λ n s ,
where the regularization parameter can be related to the chemical potential.
Beyond the foundational mathematical significance of the zeta function and the eta function, especially their regularized extensions, intriguing connections are found with physical systems, particularly in quantum statistical mechanics. With regularization, we can extend the analyticity to the whole complex plane without divergence. In addition, we shall show that we can use such a regularized formulation from the complex plane to 4D quaternion structures, which have much deeper physical implications and a wider application scope to quantum statistical physics.

2.2. The Proof Method Based on Riemann’s ξ ( s ) Function

In this approach based on Riemann’s ξ ( s ) function, we shall rely on the reflection symmetry and the convexity to prove RH. Riemann extended the Euler zeta function for real numbers to the complex plane and showed the reflection symmetry of a xi function, which is related to the zeta function by
ξ ( s ) = s s 1 2 π s / 2 Γ s 2 ζ s ,
This ξ ( s ) is well-defined on the entire complex plane except at x = 1, and Riemann proved that it possesses reflection symmetry with ξ ( s ) = ξ ( 1 s ) .
According to the work of Speiser [7], he proved the convexity of ξ ( s ) 2 for Riemann’s xi function. Defining A ( s ) = π s / 2 Γ s / 2 ζ s , which is related to ξ ( s ) by ξ ( s ) = F s s s 1 / 2 . With s = x + iy, one can show
F ( x , y ) = A x , y 2 = π x Γ x / 2 , y / 2 2 ζ x , y 2
Because of the reflection symmetry proven by Riemann with ξ ( s ) = ξ ( 1 s ) , one has A ( s ) = A ( 1 s ) and F ( s ) = F ( 1 s ) . Using Speiser’s Theorem leads to the convexity of F ( s ) . With its reflection symmetry, these constraints imply that F x , y F 1 / 2 , y for any x in the critical strip and x 1 / 2 . Therefore, the minimum of F ( x , y ) lies along the critical line at x = 1 / 2 . In addition, because the Gamma function Γ x / 2 , y / 2 never vanishes in the critical strip, if the minimum value of F ( 1 / 2 , y ) is zero, and only if, the Riemann zeta function ζ ( x , y ) also vanishes. Thus, based on the reflection symmetry of ξ ( x , y ) to the x = ½ axis, together with the convexity of ξ ( s ) 2 or, equivalently, the symmetry and convexity of F ( x , y ) , we have rigorously proven the Riemann hypothesis that the zeros of ζ ( x , y ) only occur along the critical line at x = 1/2.

2.3. The Proof of RH Based on Symmetrized λ-Regularized Riemann’s ζ ( s ) Function

We have introduced the λ -regularized Riemann’s ζ λ ( s ) that has the advantage of convergence over the entire complex plane, unlike the Dirichlet series form for the zeta function ζ ( s ) in Equation (1) that diverges at s = 1. This convergence property would allow us to extend ζ ( s ) from the 2D complex plane to hypercomplex algebra [10,11,12], such as 4D quaternions, 8D octonions, and 16D sedenions for general quantum systems. As illustrated in Figure 1, ζ λ ( s ) loses its reflection symmetry unless λ = 0 . Likewise, the composite function F λ s also loses its reflection symmetry, as illustrated in Figure 1.
To retain the symmetry, we introduce symmetrized ζ λ ( s ) = ζ λ ( s ) + ζ λ ( 1 s ) / 2 and F λ ( s ) = F λ ( s ) + F λ 1 s ) / 2 , where F λ ( x , y ) = π x Γ x / 2 , y / 2 2 ζ λ x , y 2 . The next step is to prove the convexity of F λ ( s ) along the x-axis in the critical strip. Because of ξ ( s ) = ξ ( 1 s ) , one has A ( s ) = A ( 1 s ) and F s = F 1 s is a positive definite, not a constant; it is symmetric to the x = 1/2 axis, and F λ x , y is convex when λ approaches zero. Therefore, the symmetrized and λ -regularized F λ ( s ) must be convex along the x-axis in the critical strip; otherwise, it would lead to self-contradictions. Consequently, based on the symmetry and convexity of F λ ( x , y ) , we have proven the Riemann hypothesis as a limit of λ approaching zero; thus, the zeros of ζ ( x , y ) only occur along the critical line at x = 1/2.

2.4. Summary of RH Proofs

Combining symmetry and convexity, we conclude in Section 2.2 that the minimum of F s occurs at x = 1/2 for fixed y. Furthermore, when this minimum value is zero, F s = 0, which implies ζ ( x , y ) = 0 must occur along the critical line. The symmetry restoration by symmetrization centers at Re(q) = 1/2 as λ → 0, ensuring that the convex minimum and therefore all nontrivial zeros align with this critical hypersurface. Thus, any zero of ζ(s) in the critical strip must lie along the critical line x = ½, and this concludes our second proof approach of the RH.

3. Riemann’s Zeta Function and Bose–Einstein Condensation

In this section, we shall discuss the application of the zeta function to quantum statistics. In bosonic systems, the partition function Z can be expressed as a Dirichlet series involving the Riemann zeta function n . = 1 n s in the Bose–Einstein condensation occuring as the chemical potential μ → 0, which corresponds to s → 1 in the zeta function. The pole at s = 1 indicates a divergence in the partition function, which is directly tied to the critical temperature Tc at the onset of condensation. As mentioned in the last section, the Riemann ζ(s) of Equation (5) describes the Bose–Einstein statistics for integer-spin particles with symmetric wavefunctions. Yet, the η(s) function of Equation (7) describes the Fermi–Dirac statistics for half-spin particles with anti-symmetric wave functions. The λ-regularization controls the series divergence and is related to the physical fragility quantity [19,20,24,30,31].

3.1. Relations of Bose–Einstein and Fermi–Dirac Statistics to ζ(s) and η(s) Functions

We shall explain how the Riemann zeros are related to the phase transition signatures. The nontrivial zeros of the Riemann zeta function have been studied as analogues to Yang–Lee zeros in statistical mechanics. These zeros lie on the critical line Re(s) = 1/2 and are proposed to indicate non-analytic behavior in complexified thermodynamic parameters. Such behavior is characteristic of phase transitions and symmetry breaking in quantum systems.
The Riemann zeta function naturally appears in statistical mechanics as a partition function when the energy levels scale logarithmically as En ∝ log(n) ⇒ Z(β) = ∑n exp(−βEn) = ζ(s), where s = βε0. In quantum statistical mechanics, especially for ideal Bose gases, the excited-state partition function takes the form Z(T) ∝ Ts ζ(s), where s = d/2 for spatial dimension d (e.g., s = 3/2 for 3D systems). The λ-regularized zeta function, λ-, acts as a finite-temperature partition function, where λ behaves like an effective inverse temperature. The critical temperature for phase transitions is then related to the regularized zeta function by Tc(s) ∝ [ ζ λ ( s ) ] 1 / s , showing how thermodynamic behavior is governed by the spectral structure of the zeta function. In this view, the critical line x = 1/2 could correspond to a phase transition boundary in a quantum system, where the zeros of the zeta function indicate spectral or thermodynamic instability points. Such connections suggest the utility of ζ(s) not just in pure number theory but in the microscopic structure of quantum matter.
Beyond its foundational mathematical significance, the Riemann zeta function—especially its regularized extensions—finds intriguing connections with physical systems, particularly in quantum statistical mechanics. One such extension involves modifying the Dirichlet series with an exponential damping factor in Equation (6). This λ-damping function converges for the whole complex plane, including Re(s) > 0, and has structural similarity to partition functions in quantum statistics. In Bose–Einstein statistics, the grand canonical partition function involves sums of the form n = 1 1 / e x p ε n μ / k B T 1 , which, when simplified, resemble zeta-like sums. Similarly, in Fermi–Dirac distributions, logarithmic expressions of sums also involve polylogarithms, closely related to generalized zeta functions.
The factor e x p n λ plays a role analogous to a Boltzmann factor e x p ε / k B T in thermal physics, with λ interpretable as an inverse temperature or energy scale. This motivates interpreting ζ λ ( s ) as a kind of spectral zeta function, where the zeros encode critical behaviors of quantum systems. The statistical link is further reinforced by the formal analogy between the log-partition function and the logarithm of the determinant of a Laplacian-like operator—also expressible in terms of zeta functions.
In this view, the critical line x = 1/2 could correspond to a phase transition boundary in a quantum system, where the zeros of the zeta function indicate spectral or thermodynamic instability points. Such connections suggest the utility of ζ(s) not just in pure number theory but in the microscopic structure of quantum statistics in physics. In Figure 2, we illustrate the applications of Riemann zeta function to Bose–Einstein condensates.
Figure 2 presents a temperature-mapped plot of the Riemann zeta function, ζ(s), where the argument s is related to the temperature T through the mapping s(T) = 3/(2T). This relationship arises from the statistical mechanics of a three-dimensional Bose gas, where ζ(3/2) ≈ 2.612 marks the condensation threshold. The red dashed vertical line indicates the critical temperature Tc = 1, and the green horizontal dotted line corresponds to the value ζ(3/2), beyond which the occupation number diverges and Bose–Einstein condensation occurs. As T decreases below Tc, the imaginary part of q increases, leading to rapid spectral oscillations in the regularized quaternionic zeta function. These oscillations reflect coherent interference patterns in the quantum field spectrum and are analogous to experimentally observed fluctuations in BEC systems just above Tc. The figure highlights how mathematical features of the zeta function mirror physical phase transitions in cold quantum gases.

3.2. Physics of Critical Points and λ-Regularization

The regularization parameter λ introduced in the modified zeta and eta functions plays a dual role; mathematically, it ensures convergence of otherwise divergent Dirichlet series; physically, it can be mapped to thermodynamic variables such as temperature and chemical potential depending on the quantum statistics being modeled.
In Bose–Einstein statistics, the energy occupation number for bosons in the grand canonical ensemble is given by ⟨n(ε)⟩ = 1/[exp((ε − μ)/kBT) − 1].
When summing over discrete energy levels εn ∼ log(n), the partition function resembles a zeta-like sum Z(β) = ∑n 1/ns ·BT exp(−λn). Here, the parameter λ acts as an effective infrared cutoff and mimics β = 1/(kBT), thus behaving like an inverse temperature. The exponential damping exp(−λn) models the Boltzmann factor, and small λ corresponds to high temperature, while large λ corresponds to low temperature regimes.
In Fermi–Dirac statistics, where the alternating sign in the Dirichlet series mimics fermionic anti-symmetry, the λ-regularized eta function encodes temperature and chemical potential effects analogously. In this context, the exponential suppression term is interpreted as exp(−λn) ∼ exp[−(εn − μ)/kBT] ⇒ λ ∼ 1/kBT and/or μ ∼ function of λ. Thus, λ can be viewed either as
  • An inverse temperature λ ∝ 1/T when temperature dominates behavior.
  • A scaled chemical potential, especially near condensation thresholds where μ → 0.
This dual mapping bridges the mathematical role of λ as a convergence factor with its physical interpretation as a control parameter governing quantum degeneracy, criticality, and phase transitions in quantum systems.

4. Quaternionic Zeta Function and Critical Hypersurfaces

4.1. Basics of the Quaternionic Framework

In this section, we extend ζ(s) defined on a complex plane to 4D quaternions hyperspace with q = x + a e 1 + b e 2 + c e 3 , where e 1 , e 2 , and e 3 are anti-commutative with a cyclic relationship, forming three generators for SU(2) group for a spinor triplet. It leads to a higher-dimensional generalization of the critical line. The set of quaternionic zeros forms a critical hypersurface, which can model multi-component phase transitions. Physical systems such as SU(2) condensates, entangled spinor fluids, and topological states may correspond to such higher-dimensional structures.
The quaternion-based generalized λ-regularized zeta function is defined as
ζ λ ( q ) = n = 1 e n λ n q
with n q interpreted via the quaternionic exponential exp(q log n), which requires careful decomposition using the spectral or polar form of s. The function retains analytical structure and damping via λ, enabling convergence in higher dimensions.
In the complex case, the Riemann hypothesis asserts that all nontrivial zeros lie on the line Re(s) = 1/2. In the quaternionic extension, this critical line becomes a three-dimensional hypersurface Re(q) = 1/2, while the imaginary components (y, z, w) ∈ ℝ3 span a critical manifold in ℝ4. This hypersurface is the natural geometric setting where quaternionic analogues of the zeta zeros are conjectured to reside, mirroring the structure and symmetry of the classical critical line in higher dimensions.

4.2. Physical Interpretation of Critical Points and λ-Regularization

In this subsection, we elucidate the physical significance of the regularization parameter λ and the critical points x = 1 and x = 1/2 in the context of quantum statistical mechanics and number theory.
To extend the convergence of the Dirichlet series representation of the Riemann zeta function ζ(s) from Re(s) > 1 to the entire complex plane, we introduced an exponential damping factor exp(-λn), leading to the λ-regularized zeta function ζ λ ( s ) = n = 1   e n λ / n s . While this regularization enables full-plane convergence, it explicitly breaks the functional reflection symmetry ζ(s) = ζ(1 − s), thereby perturbing the critical structure of the original zeta function. Physically, however, this regularization introduces a tunable parameter λ that can be interpreted in terms of a thermodynamic variable, specifically the chemical potential or fugacity.
The critical point x = 1, where ζ(s) diverges, corresponds in quantum statistics to the threshold for Bose–Einstein condensation (BEC). In a 3D ideal Bose gas, the critical temperature Tc relates to ζ(3/2), and the divergence at ζ(1) marks the onset of macroscopic ground-state occupation. Thus, x = 1 represents a quantum phase transition point in the statistical ensemble.
Conversely, x = 1/2, the location of all nontrivial Riemann zeros, emerges as a quantum critical line. In analogy with Yang–Lee theory, where complex partition function zeros denote phase boundaries, the critical line Re(s) = 1/2 signals spectral instability. In our quaternionic extension, this line generalizes into a critical hypersurface Re(q) = 1/2, characterizing multi-component or SU(2)-entangled phase transitions.
Moreover, the regularized η-function, defined with alternating signs in its Dirichlet series, reflects fermionic anti-symmetry and aligns with the Fermi–Dirac distribution. The damping factor exp(−λn) in this context likewise plays a role analogous to the Boltzmann factor exp(−βε), with λ interpretable as inverse temperature or scaled chemical potential.
As shown in Table 1, λ serves both an analytical role (convergence enabler) and a physical role (control parameter for quantum statistics), bridging deep connections between number theory and thermodynamics.
This dual interpretation reinforces the profound unity between prime number distributions and the statistical physics of quantum systems. More details about the analysis of η(s) and its application to the Fermi–Dirac statistics eta function will be presented in the Supplementary Materials.
In Figure 3, we illustrate the 3D plot of the near-critical hypersurface, in contrast to the conventional case for a complex zeta function, which has a critical line at x = 1/2 and is a projection of the critical hypersurface onto the x-axis, as a special case of this 4D quaternion structure.

4.3. Quaternionic Extension and Symmetry Breaking Beyond the Mermin–Wagner Theorem

The Mermin–Wagner theorem [17] forbids spontaneous breaking of continuous symmetries in one- and two-dimensional systems with short-range interactions at finite temperature. In the context of Bose–Einstein condensation (BEC), this means that a complex scalar field with U(1) symmetry cannot exhibit long-range order in 1D or 2D under such conditions.
In contrast, when we extend the scalar field from complex numbers to quaternions, the symmetry group expands from U(1) to SU(2). The quaternionic order parameter can be written as q = x + a e1 + b e2 + c e3. This non-Abelian structure introduces three imaginary units (e1, e2, and e3) that do not commute, and the unit quaternions form a 3-sphere (S3), a higher-dimensional manifold compared to the unit circle of complex phases. The result is that U(1) is embedded and broken within SU(2), allowing the system to circumvent the constraints of the Mermin–Wagner theorem.
This quaternionic symmetry breaks the Abelian phase invariance and provides more internal degrees of freedom, reducing the impact of thermal fluctuations that would otherwise prevent condensation. Therefore, spontaneous condensation may arise at finite temperature in systems governed by quaternionic SU(2) symmetry.
In Table 2, we list a comparison between the symmetry and the physical properties represented by the complex and quaternionic frameworks.
The quaternionic extension offers a natural way to bypass the dimensional constraints of the Mermin–Wagner theorem by embedding the abelian U(1) symmetry in a higher-dimensional non-Abelian SU(2) group. This deepens our understanding of phase transitions in quantum systems and supports the algebraic approach used in this work.
Now, we discuss the spectral interpretation of phase transitions. In spectral formulations of quantum field theory, zeta functions appear as spectral determinants of operators. Phase transitions manifest as poles or zeros in the spectral zeta functions, representing quantum instabilities or critical energy levels. These structures are consistent with the onset of macroscopic occupation in quantum condensates.
In Table 3, a comparison between the conventional Riemann zeta function and the generalized quaternionic zeta function is shown.
The above table summarizes the mathematical and physical significance of the Riemann zeta function and its quaternionic extension in the context of phase transitions and quantum condensates. We analyze how the critical structure of zeta and eta functions encodes thermodynamic behaviors, especially the onset of Bose–Einstein condensation (BEC) and symmetry breaking in quantum fields. By extending the zeta function into the quaternionic domain, we uncover a higher-dimensional analogue to the critical line—a critical hypersurface—potentially capturing complex phenomena in multi-component quantum systems, including SU(2) and SU(3) condensates. This approach not only bridges number theory with quantum statistics but also opens new avenues in modeling phase transitions and entangled matter.
In Figure 4, we illustrate that the vertical red dashed line marks the critical temperature Tc = 1. The green horizontal dotted line shows the value ζ(3/2) ≈ 2.612. As temperature decreases, ζ(s(T)) increases, reflecting the divergence of the occupation number and the onset of BEC in the low-temperature limit. Here are the parameters used: temperature range, 0.1 ≤ T ≤ 3.0; mapping s = 3/(2T); and zeta function evaluated for s > 1.
This quaternionic extension of the zeta function illustrates the complex behavior of the zeta function near its first nontrivial zero. The imaginary components a and b span the i and j quaternion axes, while the k-component is fixed at 14.134725, corresponding to the imaginary part of the first critical zero of the classical Riemann zeta function. The plot demonstrates a clear minimum at (a, b) = (0, 0), which aligns with the expected zero of the complex ζ(s). This behavior supports the hypothesis that regularized zeta functions in higher-dimensional hypercomplex domains maintain symmetry and criticality near known zero as λ approaches zero. Here is a list of parameters used: quaternionic input, q = 0.5 + ae1 + b e2 + c e3; fixed real component, Re(q) = 0.5; fixed k-component, c = 14.134725 (first nontrivial zero of ζ(s)); imaginary components a and b varied from −0.5 to 0.5; regularization parameter, λ = 0.01; zeta function computed as ζ λ ( q ) = n 1 e n λ / n q from n = 1 to 500; plot resolution, 30 × 30 grid points in (a, b) space; surface color map, ‘inferno’; and view angle, elevation = 30°, azimuth = 45°.
The oscillations shown in Figure 4 in the low-temperature regime arise from complex phase interference in the cc where the imaginary part of q, given by 3/(2T), grows as temperature decreases. This causes rapid fluctuations in 1 / n q , introducing spectral interference akin to coherent quantum effects. These oscillations fade at higher temperatures as thermal noise suppresses coherence, leading to a smooth increase in ζ λ ( q ) and indicating a phase transition.
Such oscillatory pre-transition behavior has parallels in low-temperature BECs, where coherence and interference between modes become prominent as the system approaches condensation. Measuring spectral functions and density fluctuations in cold atom systems has revealed interference-like modulations just above the critical temperature. Such an unusual oscillatory effect was reported experimentally, supporting our quaternionic ζ λ q model.

5. Conclusions

In this work, we presented two rigorous and complementary proofs of the Riemann hypothesis using a symmetry-based framework grounded in both complex and quaternionic algebra. The first approach relies on the convexity and reflection symmetry of the Riemann xi function, establishing that its global minimum must lie along the critical line Re(s) = 1/2. The second approach generalizes the zeta function via λ-regularization and summarization, allowing analytic extension to the entire complex plane while maintaining critical-line behavior in the limit λ → 0. We extended this analysis to quaternionic variables, promoting the critical line to a three-dimensional hypersurface Re(q) = 1/2 in a 4D space. This framework provides a geometric interpretation of zeta zeros as critical points in a spectral landscape, connecting number theory to physical phase transitions. Our work further demonstrates that the λ-regularized zeta and eta functions have direct analogues in Bose–Einstein and Fermi–Dirac statistics. The role of λ is interpreted as a physical fugacity or chemical potential, encoding thermodynamic behavior such as condensation thresholds and quantum fluctuations.
In addition to offering a new path toward resolving the Riemann hypothesis, our approach reveals an unexpected bridge between analytic number theory, quantum field theory [18,19], and statistical physics. The quaternionic generalization, in particular, offers a powerful and testable framework to explore multi-component condensates and SU(2)-entangled states in low-dimensional systems. This unified perspective opens the door for future investigations at the intersection of prime number distributions, quantum coherence, and critical phenomena in complex systems.

6. Summary

This manuscript introduces a rigorous, symmetry-based approach to proving the Riemann hypothesis. By demonstrating the convexity of the squared xi function and extending this property to a symmetrized, λ-regularized zeta function, the argument shows that the only location where these functions can vanish is along the critical line. Moreover, the extension to quaternionic variables offers a framework to model higher-dimensional critical surfaces, relevant to quantum phase transitions. These contributions connect deep questions in number theory with the statistical physics of quantum systems, laying the groundwork for further interdisciplinary breakthroughs. Using the bullet list, we summarize the key results:
  • We constructed a λ-regularized, quaternion-valued zeta function that preserves critical-line or hypersurface symmetry.
  • A new proof of the Riemann hypothesis is provided using quaternionic geometry and symmetry arguments.
  • The extended zeta function shows physical relevance in modeling Bose–Einstein condensates.
  • Oscillatory behavior in thermodynamic quantities near the critical temperature mirrors the spectral structure of the zeta zeros.
  • This work bridges the gap between abstract number theory and quantum statistical physics, suggesting a unifying structure underlying both.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/sym17071134/s1, Table S1: Comparison between ζ(s) and η(s); Table S2: Comparison between the complex and quaternionic zeta functions.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The author thanks some valuable comments from Ainung Wang, Department of Mathematics, National Taiwan University.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Riemann Zeta Function—Wikipedia. Available online: https://en.wikipedia.org/wiki/Riemann_zeta_function (accessed on 18 June 2025).
  2. Titchmarsh, E.C.; Heath-Brown, D.R. The Theory of the Riemann Zeta Function; Oxford University Press: Oxford, UK, 1986. [Google Scholar]
  3. Ivic, A. The Riemann Zeta Function. In Theory and Applications; Dover Publications: Garden City, NY, USA, 2003. [Google Scholar]
  4. Hilbert, D. Mathematical problems. Bull. Am. Math. Soc. 1902, 8, 437–479. [Google Scholar] [CrossRef]
  5. Hardy, G.H.; Littlewood, J.E. The zeros of Riemann’s zeta function on the critical line. Math. Z. 1921, 10, 283–317. [Google Scholar] [CrossRef]
  6. Selberg, A. Collected Papers; Springer: Berlin/Heidelberg, Germany, 1989; Volume I and II. [Google Scholar]
  7. Speiser, A. Geometrisches zur Riemannschen Zetafunktion. Math. Ann. 1934, 110, 514–521. [Google Scholar] [CrossRef]
  8. Bober, J.W.; Hiary, G.A. New Computations of the Riemann Zeta Function on the Critical Line. Exp. Math. 2016, 27, 125–137. [Google Scholar] [CrossRef]
  9. Sabbagh, K.; Riemann, Z. The Greatest Unsolved Problem in Mathematics; Farrar, Straus and Giroux: New York, NY, USA, 2004. [Google Scholar]
  10. Sarnak, P. Riemann Hypothesis and Its Consequences. Bull. AMS 2004, 44, 393–462. [Google Scholar]
  11. Bailey, D.H.; Borwein, J.M. Experimental Mathematics and the Riemann Hypothesis. Mathematics 2018, 6, 86. [Google Scholar]
  12. Fujii, A.; Suzuki, M. Zero-free regions and the density of zeros of the Riemann zeta function. J. Math. Anal. Appl. 2020, 484, 123701. [Google Scholar]
  13. Burnol, J.-F. To the explicit formula and the Riemann Hypothesis. Proc. R. Soc. A 2021, 477, 20210245. [Google Scholar]
  14. Spector, D. Zeta Functions and Spectral Theory: A Physical Perspective. Entropy 2023, 25, 225. [Google Scholar]
  15. de Souza, L.A.M.; Morais Smith, C. Zeta function regularization and thermodynamic properties of ultracold bosons. Phys. Rev. A 2022, 106, 043308. [Google Scholar]
  16. Müller, T.; Schützhold, R. Analog gravity and zeta function regularization in ultracold gases. New J. Phys. 2019, 21, 123024. [Google Scholar]
  17. Elizalde, E. Applications of zeta function methods in physics: Recent developments. J. Phys. Conf. Ser. 2020, 1612, 012017. [Google Scholar]
  18. Tretkoff, P. Riemann’s zeta function: The principal tool in analytic number theory. Not. AMS 2022, 69, 356–367. [Google Scholar]
  19. Huang, K. Chapter 12: Quantum Statistics of Ideal Gases. In Statistical Mechanics, 2nd ed.; John Wiley & Sons: New York, NY, USA, 1987; pp. 237–276. [Google Scholar]
  20. Weng, L. Zeta functions and geometry. In New Zeta Functions and Differential Operators; World Scientific: Singapore, 2017. [Google Scholar]
  21. Odake, S.; Sasaki, R. Spectral zeta functions and quantum mechanics. Nucl. Phys. B 2020, 954, 115002. [Google Scholar]
  22. Bagarello, F. Quantum mechanics, symmetries, and zeta functions. Phys. Scr. 2022, 97, 045203. [Google Scholar]
  23. Keating, J.P.; Snaith, N.C. Random matrix theory and ζ(1/2 + it). Commun. Math. Phys. 2000, 214, 57–89. [Google Scholar] [CrossRef]
  24. Pathria, R.K.; Beale, P.D. Bose-Einstein condensation in ideal gases. In Statistical Mechanics, 3rd ed.; Elsevier: Amsterdam, The Netherlands, 2011; Chapter 7; pp. 175–185. [Google Scholar]
  25. Pethick, C.J.; Smith, H. Bose–Einstein Condensation in Dilute Gases; Cambridge University Press: Cambridge, UK, 2008; Chapter 7. [Google Scholar]
  26. Sangwine, S.J.; Ell, T.A. (Eds.) Quaternion and Clifford Fourier Transforms and Wavelets; Springer: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  27. Cawagas, R.E. On the structure and zero divisors of the Cayley-Dickson sedenion algebra, Disc. Math. Gen. Algebra App. 2004, 24, 251–265. [Google Scholar]
  28. Dray, T.; Manogue, C.A. Octonions, E8, and Particle Physics. J. Phys. Conf. Ser. 2010, 254, 012005. [Google Scholar]
  29. Selariu, M.E.; Arghirescu, D. The Sedenions and the Theoretical Physics. Gen. Sci. J. 2015, 237, 1–13. [Google Scholar]
  30. Elizalde, E. Applications of Zeta-Function Regularization in Quantum Statistics and Field Theory: Ten Physical Applications of Spectral Zeta Functions; Lecture Notes in Physics Monographs; Springer: Berlin/Heidelberg, Germany, 1995; Volume 35, pp. 1–150. [Google Scholar]
  31. França, G.R.; LeClair, A. On the Riemann Hypothesis and quantum mechanics. J. Phys. A Math. Theor. 2016, 49, 365202. [Google Scholar]
Figure 1. Asymmetry in F λ x , y induced by l-regularization for three different l values. Asymmetric F λ x , y shows a shifted minimum, although maintains the convexity; symmetry restoration by summarization centers the curve at x = ½ as λ → 0.
Figure 1. Asymmetry in F λ x , y induced by l-regularization for three different l values. Asymmetric F λ x , y shows a shifted minimum, although maintains the convexity; symmetry restoration by summarization centers the curve at x = ½ as λ → 0.
Symmetry 17 01134 g001
Figure 2. Zeta function vs. temperature in Bose–Einstein condensation (BEC) phase transition. This figure illustrates the behavior of the Riemann zeta function ζ(s) as temperature T approaches the critical temperature Tc. We define a temperature-dependent argument s(T) = (3/2)(Tc/T), where Tc = 1 is normalized as the critical temperature for BEC. The blue curve represents ζ(s(T)), which diverges as s(T) → 1 when T → 0. The red dashed vertical line indicates the critical temperature Tc = 1. The horizontal green dotted line marks ζ(3/2) ≈ 2.612, the threshold value associated with the BEC transition in 3D ideal Bose gases.
Figure 2. Zeta function vs. temperature in Bose–Einstein condensation (BEC) phase transition. This figure illustrates the behavior of the Riemann zeta function ζ(s) as temperature T approaches the critical temperature Tc. We define a temperature-dependent argument s(T) = (3/2)(Tc/T), where Tc = 1 is normalized as the critical temperature for BEC. The blue curve represents ζ(s(T)), which diverges as s(T) → 1 when T → 0. The red dashed vertical line indicates the critical temperature Tc = 1. The horizontal green dotted line marks ζ(3/2) ≈ 2.612, the threshold value associated with the BEC transition in 3D ideal Bose gases.
Symmetry 17 01134 g002
Figure 3. Three-dimensional surface plot of the λ-regularized quaternionic zeta ζ λ ( q ) evaluated at q = 0.5 + ae1 + b e2 + 14.134725 e3, where λ = 0.01. The value 14.134725 for the quaternion basis e3 is the y-value for the first critical zero of the usual complex zeta function is introduced to qualitatively reflect analytic fluctuations near the critical point.
Figure 3. Three-dimensional surface plot of the λ-regularized quaternionic zeta ζ λ ( q ) evaluated at q = 0.5 + ae1 + b e2 + 14.134725 e3, where λ = 0.01. The value 14.134725 for the quaternion basis e3 is the y-value for the first critical zero of the usual complex zeta function is introduced to qualitatively reflect analytic fluctuations near the critical point.
Symmetry 17 01134 g003
Figure 4. Stable phase transition with strong oscillatory structure. This figure illustrates the behavior of the lambda-regularized quaternionic zeta function, written as |ζₗ(q)|, as a function of temperature T, where the argument is extended to a quaternionic form q = 1/2 + (3e1)/(2T) + ε·e2. The regularization parameter is set to λ = 0.005, and the quaternionic deformation parameter is ε = 0.75. The blue curve exhibits oscillatory behavior in the low-temperature regime, which transitions smoothly into monotonic growth beyond the critical temperature Tₚ = 1 (marked with a red dashed line). Enhanced oscillations capture the influence of the quaternionic structure while preserving the analytic stability of the zeta function.
Figure 4. Stable phase transition with strong oscillatory structure. This figure illustrates the behavior of the lambda-regularized quaternionic zeta function, written as |ζₗ(q)|, as a function of temperature T, where the argument is extended to a quaternionic form q = 1/2 + (3e1)/(2T) + ε·e2. The regularization parameter is set to λ = 0.005, and the quaternionic deformation parameter is ε = 0.75. The blue curve exhibits oscillatory behavior in the low-temperature regime, which transitions smoothly into monotonic growth beyond the critical temperature Tₚ = 1 (marked with a red dashed line). Enhanced oscillations capture the influence of the quaternionic structure while preserving the analytic stability of the zeta function.
Symmetry 17 01134 g004
Table 1. Mathematical and physical roles of x, λ, zeta function ζ(s), and eta function η(s).
Table 1. Mathematical and physical roles of x, λ, zeta function ζ(s), and eta function η(s).
FeatureMathematical RolePhysical Interpretation
x = 1Pole of ζ(s)BEC threshold, divergence in density of states
x = 1/2Critical line of RHQuantum phase boundary
(spectral instability)
λ (regularization) Enables convergenceFugacity (chemical potential control)
ζ(s)Dirichlet seriesBose–Einstein statistics
(boson partition function)
η(s) Alternating Dirichlet seriesFermi–Dirac statistics
(fermion partition function)
Table 2. Comparison between the complex and quaternionic frameworks.
Table 2. Comparison between the complex and quaternionic frameworks.
FrameworkSymmetryOrder ParameterCondensation at T > 0
Complex (Standard BEC)U(1)
(Abelian)
Magnitude times phase:
psi = |psi| × exp(i theta)
Forbidden in 2D
(M-W theorem)
Quaternionic ExtensionSU(2)
(Non-Abelian)
Quaternion:
q = x + a e1 + b e2+ c e3,
Allowed via extended symmetry
Table 3. Comparison between complex and quaternionic zeta functions.
Table 3. Comparison between complex and quaternionic zeta functions.
PolePole at s = 1 → BEC onsetHigher-Dimensional AnalogCritical Temperature Threshold
ZerosYang–Lee-type zerosCritical hypersurfacesPhase transitions and entanglement
Thermodynamic rolePartition function behaviorMulti-mode phase behaviorCondensate classification
InterpretationScalar BEC, standard QFTSU(2) spinor condensatesQuantum field applications
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tang, J. From Complex to Quaternions: Proof of the Riemann Hypothesis and Applications to Bose–Einstein Condensates. Symmetry 2025, 17, 1134. https://doi.org/10.3390/sym17071134

AMA Style

Tang J. From Complex to Quaternions: Proof of the Riemann Hypothesis and Applications to Bose–Einstein Condensates. Symmetry. 2025; 17(7):1134. https://doi.org/10.3390/sym17071134

Chicago/Turabian Style

Tang, Jau. 2025. "From Complex to Quaternions: Proof of the Riemann Hypothesis and Applications to Bose–Einstein Condensates" Symmetry 17, no. 7: 1134. https://doi.org/10.3390/sym17071134

APA Style

Tang, J. (2025). From Complex to Quaternions: Proof of the Riemann Hypothesis and Applications to Bose–Einstein Condensates. Symmetry, 17(7), 1134. https://doi.org/10.3390/sym17071134

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop