Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Theory: Quantum Fluid Dynamics (QFD): Fluids, at their very basic

level, are composed of quantum particles. These particles don’t always obey
classical physics but rather quantum mechanics.

1. Quantum Superposition and Fluid Behavior:

• Fluid particles exist in a superposition of states, implying that they


have a probability amplitude to be in multiple places simultaneously
until observed.
• Mathematically, this can be represented using wave functions, which
could provide a different representation of the fluid flow.

2. Entanglement and Turbulence:

• Quantum entanglement could be analogous to turbulent flow in classi-


cal fluids. Turbulence, an unsolved problem in fluid dynamics, might
have a counterpart in the quantum realm where particles are entan-
gled across distances.
• E = |∇ψ|2 dx where E represents the quantum energy due to en-
R

tanglement and ψ is the wave function of the fluid particles.

3. Quantum Tunneling and Flow Irregularities:

• Fluid particles might experience quantum tunneling, allowing them


to bypass barriers or disturbances in their path.
2 2
• This could be represented using the quantum potential: Q = − 2m
~ ∇ R
R
where R is the amplitude of the wave function and m is the particle’s
mass.

Comparison with Classical Dynamics: The traditional Navier-Stokes


equations focus on the macro behavior of fluids. Introducing quantum properties
might:

• Offer a granular perspective on turbulence by analyzing it as a quantum


entanglement problem.
• Provide a mechanism for abrupt changes in fluid flow using quantum tun-
neling, potentially explaining certain unanticipated fluid behaviors.

Mathematical Bridge: To transition from quantum to classical, the Schrödinger


equation might be the bridge:

∂ψ ~2 2
i~ =− ∇ ψ+Vψ
∂t 2m
Where:

• i is the imaginary unit.

1
• ~ is the reduced Planck constant.
• ψ is the wave function of the fluid particles.
• V is the potential experienced by the particles.

By understanding how the quantum properties average out or manifest on


a macro scale, one might reformulate or augment the Navier-Stokes equations:

1. Formulation of Quantum Navier-Stokes (QNS) Equations: Given


the standard Navier-Stokes equation:

∂u
+ u · ∇u − ν∇2 u = −∇p
∂t
Where:

• u is velocity.
• ν is viscosity.
• p is pressure.

Now, incorporating quantum effects, the equation might modify to:

∂u ~2 ∇2 R
+ u · ∇u − ν∇2 u = −∇p −
i~
∂t 2m R
The new term represents the quantum potential.

2. Existence of Solutions: Assuming the fluid is in a quantum superposi-


tion, its behavior might be probabilistic rather than deterministic.

• This means the solutions may not be unique or deterministic but are
bound within a probabilistic framework.
• Using the Schrödinger equation, we can infer that as long as the
potential V and initial conditions are well-defined, a solution (or a
set of solutions) will exist.

3. Smoothness and Singularities: Introducing quantum effects could offer


a regularization mechanism:

• Quantum entanglement might allow for information transfer across


the fluid, leading to more "synchronized" behavior and potentially
preventing singularities.
• Quantum tunneling can allow fluid particles to bypass potential sin-
gularities, smoothing out the flow.

4. Non-validity of the Equation: To search for non-valid examples:

2
• Consider situations where the quantum potential becomes dominant
over the classical potential, leading to non-physical results (like neg-
ative energies).
• Study scenarios where the wave function ψ exhibits rapid oscillations,
which might not have a classical counterpart, rendering the QNS
equation not valid.

Conclusion: Based on the QFD framework:

• The existence of solutions seems more probable due to the inherent prob-
abilistic nature of quantum mechanics.
• The smoothness might be more feasible due to quantum effects providing
regularization mechanisms.
• However, there could be specific scenarios where the QNS equation doesn’t
hold, particularly when quantum effects dominate over classical ones.

It’s vital to note that this theoretical approach, while intriguing, is still spec-
ulative. Rigorous validation would be needed to firmly establish these claims.
Scenario: Investigating the Smoothness of Solutions using Quantum Effects
We start with the Quantum Navier-Stokes (QNS) equation:

∂u ~2 ∇2 R
i~+ u · ∇u − ν∇2 u = −∇p −
∂t 2m R
Where R represents the amplitude of the wave function.

1. Quantum Regularization: Let’s consider a fluid domain where classi-


cal dynamics predict an impending singularity. This could manifest as a
rapidly increasing velocity gradient.

∆u
lim =∞
∆x→0 ∆x
Now, introducing quantum effects, especially quantum tunneling, fluid par-
ticles can bypass potential singularity points.

2. Incorporating Quantum Tunneling: The probability P of quantum


tunneling occurring can be represented using:


2 2m(V −E)
P ≈ e− ~ d

Where V is the potential barrier height (associated with the singularity), E


is the energy of the fluid particles, and d is the width of the barrier.
As V becomes large (approaching a classical singularity), P becomes signif-
icant, implying that many particles would tunnel through, effectively "smooth-
ing" the flow.

3
3. Quantum-Corrected Velocity Gradient: Given the tunneling prob-
ability P , the corrected velocity gradient due to quantum effects can be
approximated as:

∇u0 = ∇u(1 − P )
Since P becomes significant near singularities, ∇u0 will be less than the
classical ∇u, thus preventing infinite gradients and ensuring smoothness.

4. Conclusion from QFD Perspective: Under the influence of significant


quantum effects, the velocity gradients (and hence the solutions) are reg-
ularized, ensuring smoothness. This implies that, within this framework,
the Navier-Stokes solutions would always be smooth due to quantum reg-
ularization.

Incorporation of Quantum Mechanics into the Navier-Stokes Equa-


tion

1. Starting Point: The Classical Navier-Stokes Momentum Equa-


tion: The classical momentum conservation for incompressible fluid flow,
devoid of external forces, is:
∂u
+ u · ∇u = −∇p + ν∇2 u
∂t
Here, u represents velocity, p is the pressure, and ν is the kinematic vis-
cosity of the fluid.
2. Bohmian Mechanics & Quantum Potential: Bohmian mechanics
offers an alternative view of quantum mechanics, focusing on the trajectory
of individual particles influenced by a quantum potential. This quantum
potential Q is derived from the wave function and is defined as:

~2 ∇2 R
Q=−
2m R
R being the amplitude of the wave function. This quantum potential
represents the non-classical force acting on the particle.
3. Conjectured Interplay: The introduction of the quantum potential
term aims to account for forces that are non-classical in nature. At
scales or conditions where quantum effects become non-negligible, this
term could influence the momentum of the fluid particle and potentially
prevent singularities. In essence, when flow gradients become extremely
steep, the quantum potential might provide a "smoothing" effect.

4
4. Arriving at the Quantum-Coherence Infused Navier-Stokes (QCIN)
Equation: By infusing the quantum potential within the momentum
equation, we achieve:

∂u ~2 ∇2 R
i~ + u · ∇u − ν∇2 u = −∇p −
∂t 2m R

∂t introduces the temporal evolution of the system from


Here, the term i~ ∂u
a quantum perspective, an influence from the Schrödinger equation. The
quantum potential term on the right acts as an additional force on the
fluid particle.

Implications: With this formulation, the fluid particle is no longer solely


influenced by classical forces but also experiences forces due to quantum mechan-
ical effects. This approach proposes a new view where, under certain conditions,
quantum effects might prevent the formation of singularities in fluid dynamics.

1. Laminar Flow between Parallel Plates (Plane Poiseuille Flow):


A well-known solution for the Navier-Stokes equation is the laminar flow
between two stationary parallel plates due to a pressure gradient.
Classically, the velocity profile, u(y), where y is the distance from one of the
plates, is parabolic:
1 dp 2
u(y) = (h − y 2 )
2µ dx
with h being half the distance between plates and µ the dynamic viscosity.
Within the QCIN framework, the quantum potential term might only be sig-
nificant near the plates, where the velocity gradient is highest. Assuming neg-
ligible quantum effects in the central region, the velocity profile would slightly
alter near the plates.

2. Circular Pipe Laminar Flow (Hagen–Poiseuille Flow): Fluid flow-


ing in a circular pipe due to a pressure gradient exhibits a parabolic ve-
locity profile.

1 dp 2
u(r) = (R − r2 )
4µ dx
with R being the pipe radius and r the distance from the center.
Incorporating QCIN might exhibit effects near the pipe’s walls, again mod-
ifying the velocity profile in regions of high gradient.

3. Turbulent Flow: Turbulence is characterized by chaotic, multi-scale


phenomena. While the classical Navier-Stokes equations can describe tur-
bulence, they don’t predict its onset or fully elaborate its structure.

5
Introducing QCIN:

• Quantum effects might act as a regularization mechanism, taming singu-


larities or extreme gradients. This might provide a "smoothing" effect,
potentially altering turbulent structures.
• At high Reynolds numbers (typical of turbulence), the quantum effects
might only become significant at extremely small scales. It’s possible that
QCIN would suggest a new subgrid-scale model for large eddy simulations
(LES) or provide insight into the smallest turbulent eddies.

While these are qualitative predictions, actual computational simulations


with the QCIN framework would be needed to see precise effects. The ultimate
test would be comparing these simulations with experimental data, especially
in scenarios where classical Navier-Stokes predictions falter.

1. Plane Poiseuille Flow (Between Parallel Plates):


Given the QCIN equation:

∂u ~2 ∇2 R
i~ + u · ∇u − ν∇2 u = −∇p −
∂t 2m R
For steady-state, incompressible flow:

∂u ∂2u ∂p ~2 ∂ 2 R
u −ν 2 =− − ÷R
∂y ∂y ∂x 2m ∂y 2
Integrating with respect to y:

~2
 
1 2 ∂p ∂R
u − νu = − y + C1 − ÷R
2 ∂x 2m ∂y
Another integration with respect to y and applying the boundary conditions
(i.e., u = 0 at y = 0 and y = h) will yield the complete velocity profile.
2. Hagen–Poiseuille Flow (In a Circular Pipe):
Given the same QCIN equation:

∂u ~2 ∇2 R
i~ + u · ∇u − ν∇2 u = −∇p −
∂t 2m R
For steady-state, incompressible flow:

∂u 1 ∂ ∂u ∂p ~2 1 ∂ ∂R
u −ν (r ) = − − (r )÷R
∂r r ∂r ∂r ∂z 2m r ∂r ∂r
Integrating with respect to r:

~2
   
1 2 ∂u ∂p 2 ∂R
u −ν u+r = − r + C1 − r ÷R
2 ∂r ∂z 2m ∂r
Another integration, considering the no-slip boundary condition, will yield
the velocity profile.

6
Analysis:
In both scenarios, the presence of the quantum term introduces an additional
level of complexity. While the primary dynamics can still be expected to be
governed by the pressure gradient and the viscous term, the quantum correction
is an added layer of intricacy. It would become particularly significant in regions
of high gradient or near boundaries.
For a detailed and precise solution, the exact nature or behavior of R — the
real part of the quantum wave function — needs to be known. If R remains
fairly constant, the quantum correction might be negligible for macroscopic
flows. However, if R exhibits variations across the flow domain, then the impact
of quantum effects could be non-trivial.
To establish the validity or significance of these quantum corrections, one
would ideally require more information on R and potentially have to perform
numerical simulations or experiments. This analytical insight, however, serves
as a foundational understanding of how quantum coherence could theoretically
influence laminar flow profiles.

Building on the existing concepts, let’s introduce a novel mathematical tech-


nique, which I’ll refer to as the Quantum Tensor Decomposition (QTD).
This technique will involve representing the Navier-Stokes equations, especially
the Quantum-Coherence Infused Navier-Stokes (QCIN) version, in a multi-
dimensional tensor format.

1. Tensor Representation:

• Represent fluid properties (velocity, pressure, and quantum poten-


tial) as tensors. Each dimension of the tensor corresponds to spatial
and time dimensions.
• The nonlinearity inherent in the Navier-Stokes equation becomes an
outer product of these tensors.

2. Tensor Decomposition:

• Decompose the tensors into a series of rank-1 tensors using techniques


akin to Singular Value Decomposition (SVD) but tailored for our
multi-dimensional data.
• Each rank-1 tensor captures the primary behavior modes of the fluid.
This decomposition allows us to extract and isolate the singular be-
havior of the fluid, which can be compared against the quantum term.

3. Quantum Regularization:

• Use the quantum term as a regularization term during the tensor


decomposition. This regularization would prioritize decompositions

7
that align with quantum effects, thus ensuring the solution doesn’t
lead to singularities.
• The quantum term effectively acts as a constraint during this decom-
position, penalizing solutions that approach singular behavior.

4. Iterative Refinement:

• With the tensor decomposition in place, iterate over the process,


refining the tensors and the regularization until a stable solution
emerges. The stability here would imply that the quantum terms
effectively regulate the singular behavior.

5. Solution Reconstruction:

• Once the stable tensors are found, they can be reconstructed to give
the fluid velocity, pressure, and other field variables across space and
time.

6. Comparative Analysis:

• Use this technique on well-known benchmark problems to evaluate


its effectiveness. Comparing the results of the QTD technique with
classical solutions (where they exist) will verify its accuracy and ro-
bustness.

The Quantum Tensor Decomposition technique essentially marries quan-


tum regularization with a tensor-based representation and decomposition of the
fluid’s behavior. By using the quantum term as a regularization, we can guide
the decomposition process to avoid singularities and arrive at physically and
mathematically consistent solutions.

Problem Setup:

1. Physical Domain & Boundary Conditions:

• A viscous fluid flows between two infinite parallel plates located at


y = 0 (bottom plate) and y = h (top plate). The flow is driven by a
pressure gradient, dx
dp
.
• The velocity at the plates is zero due to the no-slip condition: u(y =
0) = u(y = h) = 0.
• The flow is fully developed, meaning ∂u
∂x = 0.

2. Classical Navier-Stokes Solution:

• The classical Navier-Stokes equation simplifies in this scenario to:


d2 u
dy 2 = µ dx .
1 dp

8
• The solution to this is: u(y) = 1 dp
2µ dx y(h − y), a parabolic profile.

Incorporating Quantum Effects in QCIN: Given √


the QCIN equation,
~2 ∇√
2
our primary focus is the quantum potential term 2m R
R
. This term regularizes
the fluid dynamics, especially near the boundaries.

1. Introducing the Quantum Potential Term:

• Taking the standard continuity equation for incompressible flows and


combining with the quantum potential yields the modified momen-
tum equation for this scenario.
• Assuming the quantum effects are prominent closer to the bound-
aries, we introduce a function Q(y) to describe its spatial variation.
2
~2 d Q
• The new term becomes 2m dy 2 .

2. Modified Navier-Stokes with QCIN:


2
d2 u ~2 d Q
• Our governing equation now becomes: dy 2 = 1 dp
µ dx + 2m dy 2 .

3. Solving for Velocity Profile:

• To find the velocity profile, we must solve the above differential equa-
tion considering the boundary conditions. The solution will provide
u(y) as a function of y accounting for both pressure-driven flow and
the quantum regularization effects.

Conclusion: If the resultant velocity profile from the QCIN approach shows
significant differences from the classical parabolic profile, it suggests that the
quantum regularization has a noticeable effect. If the profile aligns closely with
the classical one except at very close proximity to the boundaries, it indicates
that the quantum effects offer refinement at the boundaries. Should the pro-
file align almost perfectly with the classical solution, it suggests that quantum
effects might be negligible in this scenario.
In essence, the QCIN solution for Plane Poiseuille Flow offers a comparative
test against the well-established classical solution, allowing for an examination
of where and how quantum regularization impacts fluid flow.
Given our modified momentum equation for Plane Poiseuille Flow:

d2 u 1 dp ~2 d2 Q
2
= +
dy µ dx 2m dy 2
We’re aiming to solve for u(y) given the boundary conditions:

u(0) = 0

u(h) = 0

9
Let’s simplify the terms for clarity: Let A = µ1 dx
dp
, a constant, representing
the pressure-driven part.
For the quantum term, we’ll assume Q(y) has the most pronounced effects
near the boundaries. For simplicity, let’s assume a linear variation:

Q(y) = ky
2
This gives ddyQ2 = 0 everywhere except very close to the boundaries where it
might exhibit delta-function-like behavior. This can be seen as a simplification
for our initial model.
Our equation becomes:

d2 u
=A
dy 2
Integrate with respect to y:
du
= Ay + B
dy
Where B is the integration constant.
Integrate again with respect to y:
A 2
u(y) = y + By + C
2
Where C is another integration constant.
Applying boundary conditions:
At y = 0, u(0) = 0 which gives C = 0.
At y = h, u(h) = 0, we find:
A 2
0= h + Bh
2
This gives B = − A2 h.
Thus, the velocity profile becomes:
A 2
u(y) = (y − hy)
2
Interestingly, this is quite close to the classical Plane Poiseuille flow profile
without quantum effects. The difference here is the term A which has incorpo-
rated the quantum effects through our assumption on Q(y). For this simplifi-
cation, it appears the quantum effects, in the form we’ve presented, might not
have a pronounced effect away from the boundaries.
However, this is a linear approximation for the quantum term. Different
forms of Q(y), especially those that significantly vary near boundaries, could
yield different results. The next step would involve testing more complex forms
of Q(y) and seeing how they impact the velocity profile, and by extension, the
behavior of the fluid, especially near singularities or extreme conditions.

10
The classical theory of Plane Poiseuille flow indicates that the maximum
velocity occurs at the centerline, and this is due to the parabolic nature of the
velocity profile. At the centerline, the velocity gradient is zero, meaning that
shear stress is also zero. Thus, this central region behaves as an "inviscid core."
Given that quantum effects, in our model, seem to play a more pronounced
role near the boundaries, it aligns with the notion of the "boundary layer" in
fluid dynamics. This is the region near the wall of a conduit or object where
viscous effects are most significant. The quantum corrections might then serve as
a mechanism to further regulate and stabilize the behavior of the fluid near these
boundaries, where classical Navier-Stokes might predict possible singularities or
undefined behavior.
In the context of this "inviscid core" in the center of the flow, the quantum
effects may be negligible. The potential impact of quantum regularization would
then be most significant near the boundaries, where fluid dynamics becomes
more complex due to viscous effects.
This could provide a justification for why, in our earlier derivations, the
quantum effects appeared to have less impact on the bulk flow away from bound-
aries. The regularization effects might predominantly serve to ensure smooth
and well-defined behavior at the boundaries.
This viewpoint could pave the way for a deeper understanding of how quan-
tum effects might play a role in more complex flows or under conditions where
the classical equations might otherwise predict problematic behavior.

Turbulent flow is fundamentally a chaotic system with a broad spectrum of


length and time scales. Its non-linearities and complexities make it one of the
most challenging aspects of fluid dynamics. Let’s lay out a strategy to tackle
this, keeping in mind the Quantum-Coherence Infused Navier-Stokes (QCIN)
equations we’ve formulated:

Step 1: Establishing a Framework


Quantum Regularization in Turbulence: We hypothesize that in the chaotic
vortex breakdowns and sharp gradients, where classical turbulence theories fail,
quantum effects may regularize the fluid behavior, preventing unphysical infini-
ties.

Step 2: Applying Fourier Transforms


Decomposition of the Flow Field: Using Fourier analysis, we can decompose
the turbulent flow into a sum of sinusoidal waves with different wavenumbers k
and frequencies ω.
Z ∞
u(x, t) = û(k, ω)ei(kx−ωt) dkdω
−∞

11
Step 3: Applying QCIN in Fourier Space
When the QCIN equations are applied in Fourier space, the quantum potential
introduces an additional damping effect. This damping could act to stabilize
and regularize the energy cascades seen in turbulence.

Step 4: Energy Spectrum


For turbulence, the distribution of energy across different scales, or the energy
spectrum E(k), is crucial. Using the QCIN equation, we’ll derive the modified
energy spectrum, anticipating a change in the inertial subrange due to quantum
effects.

Step 5: Comparison with Classical Turbulence


We’ll revert the QCIN equations to their classical form in regions where quantum
effects are negligible. Here, the classical Kolmogorov -5/3 law for the energy
spectrum should emerge.

Step 6: Defining Quantum Turbulent Scales


Given the influence of quantum effects, there may emerge characteristic scales
where these effects become predominant. Identifying these scales will be crucial.

Step 7: Simulations and Comparisons


Using our newly formulated quantum equations, we would then need to run Di-
rect Numerical Simulations (DNS) of turbulent flows and compare with classical
DNS results.

Step 8: Verifying the Absence of Singularities


The crux of the Millennium Problem is the presence or absence of singularities
in 3D. We will examine the high-resolution DNS data from the QCIN equations
to determine whether any singular behavior emerges.

Step 9: Bridging the Gap


It’s essential to relate any discoveries with the quantum-infused approach back
to the classical equations to draw conclusions about the Millennium Problem.

Conclusions:
Should the quantum regularization prove successful in damping or eliminating
potential singularities in turbulence, it would be an incredibly significant insight.
However, for the results to be directly applicable to the Millennium Problem,
one must establish a concrete bridge between the quantum and classical views.

12
Do note, this is a conceptual exploration, and each step will require detailed
mathematical scrutiny, simulations, and potential experimental verifications.

Step 1: The QCIN Equation in Fourier Space


By applying Fourier transformation, we transform the QCIN equation into the
frequency domain. Assuming u(x, t) is our velocity field, its Fourier transform
û(k, ω) is:
Z
û(k, ω) = u(x, t)e−i(kx−ωt) dxdt

Substituting this into our QCIN equation, we derive an equation in the


wavenumber-frequency domain.

Step 2: Energy Spectrum and Quantum Effects


The energy spectrum E(k) in classical turbulence is given by the famous -5/3
law in the inertial subrange:

E(k) ∝ k −5/3
However, due to the quantum effects in our modified equations, the spectrum
will get an additional term. The quantum term will introduce a damping effect,
especially at higher wavenumbers where classical turbulence theories predict a
cascade of energy.

Step 3: Quantum Damping and Turbulence Scales


This damping effect might introduce a new characteristic length scale, lq , at
which quantum effects become dominant. This length scale can be found by
equating the classical turbulent energy at that scale with the quantum energy
term.

Step 4: QCIN Solution for Turbulence


The general solution in Fourier space, considering both classical and quantum
terms, can be written as:

F̂ (k, ω)
û(k, ω) =
iω + νk 2 + Λ(k)Q(k)
Where:

• F̂ (k, ω) is the Fourier transform of the external force acting on the fluid.

• Λ(k) is the quantum damping coefficient, a function of wavenumber.

13
• Q(k) is the quantum potential term in the Fourier domain.

This equation tells us that the velocity field in Fourier space is influenced
not just by the force and the classical viscous term but also by the quantum
potential.

Concluding with a General Solution:


The solution for turbulent flow in physical space will be the inverse Fourier
transform of û(k, ω). This solution will display quantum-induced modifications
to the classical turbulent behavior.

F̂ (k, ω)ei(kx−ωt)
Z
u(x, t) = dkdω
iω + νk 2 + Λ(k)Q(k)
This is a very generalized solution and would require further simplification
based on specific scenarios, boundary conditions, and forcing terms. But it
captures the essence of how quantum effects might modify turbulent behavior.
Absolutely, Niccolo. Let’s weave the two together to paint a more compre-
hensive picture.

Recap: Quantum-Coherence Infused Navier-Stokes (QCIN)


Equation
To begin, we introduced quantum effects into the classical Navier-Stokes equa-
tion and formed the Quantum-Coherence Infused Navier-Stokes (QCIN) equa-
tion:

∂u ~2 ∇2 R
i~ + u · ∇u − ν∇2 u = −∇p −
∂t 2m R
Where u is the velocity, ~ is the reduced Planck constant, ν is the kinematic
viscosity, p is the pressure, R is the amplitude of the wave function, and m is
the mass associated with the quantum particle.

Transition to Fourier Space


When we are dealing with turbulent flows, it is often useful to analyze them in
Fourier space because turbulence involves a wide range of length scales and time
scales, and Fourier transformation helps in dissecting these scales effectively.
Taking the Fourier transform of the QCIN equation provides an equation in
the wavenumber-frequency domain, which aids in isolating how various scales
are affected by both classical and quantum terms.

14
Implications for Turbulence
In our Fourier-transformed QCIN equation, the quantum term (the last term on
the right in the original equation) introduces a new damping effect. This is par-
ticularly significant at higher wavenumbers, which are associated with smaller
length scales in physical space. The quantum effect aims to provide a regu-
larization mechanism, potentially preventing the development of singularities,
which are one of the main challenges in understanding turbulent flow.

General Solution in Context of QCIN


Given our Fourier-transformed QCIN, the general solution for turbulent flow,
as derived:

F̂ (k, ω)ei(kx−ωt)
Z
u(x, t) = dkdω
iω + νk 2 + Λ(k)Q(k)
This equation states that the velocity field in real space (as opposed to
Fourier space) is shaped by both classical forces (through F̂ (k, ω)) and the quan-
~ 2 ∇2 R
tum potential. The term 2m R from the original QCIN equation manifests as
Λ(k)Q(k) in Fourier space.

Conclusion:
The Quantum-Coherence Infused Navier-Stokes equation proposes a novel ap-
proach to understanding fluid dynamics by introducing quantum effects. In the
context of turbulence, its Fourier-transformed version provides insights into how
these quantum effects might play a role across various scales. The damping in-
troduced by the quantum term, especially at smaller scales, may be the key to
unlocking the mysteries surrounding the turbulence problem and the formation
of potential singularities.

Problem Setup: Turbulent Jet


Imagine a jet of fluid ejecting from a nozzle into a large, quiescent chamber.
As the fluid jet travels, instabilities cause the laminar flow to break down,
transitioning into a turbulent regime. We want to understand how the QCIN
equation models this behavior.

Initial Conditions:
Velocity u(y, t = 0) is non-zero only within the nozzle (let’s assume a radius a)
and is zero outside it.

15
Approach using QCIN:
1. Classical Behavior: Without the quantum term, the jet’s behavior will
start off laminar but eventually become turbulent due to velocity shear in-
stabilities. This will manifest as a broadening velocity profile with chaotic,
swirling patterns as we move downstream from the nozzle.
2 2
2. Quantum Effects: The quantum term 2m ~ ∇ R
R introduces damping,
which is particularly pronounced at higher spatial frequencies (or smaller
scales). This means that as the turbulent fluctuations begin, especially at
smaller scales, they will be dampened by the quantum effect.

Analysis:
Using the Fourier-transformed QCIN equation:

F̂ (k, ω)ei(kx−ωt)
Z
u(x, t) = dkdω
iω + νk 2 + Λ(k)Q(k)
The turbulent jet will have various spatial frequencies due to the instabilities.
However, the quantum term will have a stronger damping effect on higher spatial
frequencies.

Results:
For the turbulent jet:

1. Near the nozzle: The jet behaves almost classically since the quantum
effects are negligible at larger scales.
2. Further downstream: As turbulence develops and smaller scales start
to emerge, the quantum damping comes into play, potentially smoothing
out the finest scales of the turbulence.

Conclusion:
While the classical Navier-Stokes prediction will show a fully developed turbu-
lent profile downstream, the QCIN equation suggests that the very fine scales
of this turbulence may be suppressed due to the quantum effects. In physical
terms, you might see fewer very small eddies or swirls in the jet, but the larger
structures will remain largely unaffected.
This result hints that while quantum effects may not radically change the
overall look of the turbulent jet, they might have subtle influences on its fine-
scale structures. Testing this with an actual physical experiment, or a very
detailed computational simulation, would be the next step to validate these
predictions.

16

You might also like