Scallop theorem

Last updated
A scallop swims by slowly opening its two halves, then quickly closing them. This is a successful swimming method because the inertial force of the surrounding water dominates over the viscous force. In a low Reynolds number environment, a scallop would only oscillate in place. Purcell's swimming scallop with reciprocal motion.png
A scallop swims by slowly opening its two halves, then quickly closing them. This is a successful swimming method because the inertial force of the surrounding water dominates over the viscous force. In a low Reynolds number environment, a scallop would only oscillate in place.

In physics, the scallop theorem states that a swimmer that performs a reciprocal motion cannot achieve net displacement in a low-Reynolds number Newtonian fluid environment, i.e. a fluid that is highly viscous. Such a swimmer deforms its body into a particular shape through a sequence of motions and then reverts to the original shape by going through the sequence in reverse. It does not matter how fast or slow the swimmer executes the sequence. At low Reynolds number, time or inertia does not come into play, and the swimming motion is purely determined by the sequence of shapes that the swimmer assumes.

Contents

Edward Mills Purcell stated this theorem in his 1977 paper Life at Low Reynolds Number explaining physical principles of aquatic locomotion. [1] The theorem is named for the motion of a scallop which opens and closes a simple hinge during one period. Such motion is not sufficient to create migration at low Reynolds numbers. The scallop is an example of a body with one degree of freedom to use for motion. Bodies with a single degree of freedom deform in a reciprocal manner and subsequently, bodies with one degree of freedom do not achieve locomotion in a highly viscous environment.

Background

Animation of a 3-sphere swimmer. It has one degree of freedom where the left arm extends and retracts. In low Reynolds number environments, this leads to no net displacement of the whole body as the arm completes a cycle of extension and retraction. NoSwim.gif
Animation of a 3-sphere swimmer. It has one degree of freedom where the left arm extends and retracts. In low Reynolds number environments, this leads to no net displacement of the whole body as the arm completes a cycle of extension and retraction.

The scallop theorem is a consequence of the subsequent forces applied to the organism as it swims from the surrounding fluid. For an incompressible Newtonian fluid with density and dynamic viscosity , the flow satisfies the Navier–Stokes equations:

where denotes the velocity of the fluid. However, at the low Reynolds number limit, the inertial terms of the Navier-Stokes equations on the left-hand side tend to zero. This is made more apparent by nondimensionalizing the Navier–Stokes equations. By defining a characteristic velocity and length, and , we can cast our variables to dimensionless form:

where the dimensionless pressure is appropriately scaled for flow with significant viscous effects. Plugging these quantities into the Navier-Stokes equations gives us:

And by rearranging terms, we arrive at a dimensionless form:

where is the Reynolds number. In the low Reynolds number limit (as ), the LHS tends to zero and we arrive at a dimensionless form of Stokes equations. Redimensionalizing yields:

Statement

The consequences of having no inertial terms at low Reynolds number are:

In particular, for a swimmer moving in the low Reynolds number regime, its motion satisfies:

Proof by scaling

This is closer in spirit to the proof sketch given by Purcell. [1] The key result is to show that a swimmer in a Stokes fluid does not depend on time. That is, a one cannot detect if a movie of a swimmer motion is slowed down, sped up, or reversed. The other results then are simple corollaries.

The stress tensor of the fluid is .

Let be a nonzero real constant. Suppose we have a swimming motion, then we can do the following scaling:

and obtain another solution to the Stokes equation. That is, if we scale hydrostatic pressure, flow-velocity, and stress tensor all by , we still obtain a solution to the Stokes equation.

Since the motion is in the low Reynolds number regime, inertial forces are negligible, and the instantaneous total force and torque on the swimmer must both balance to zero. Since the instantaneous total force and torque on the swimmer is computed by integrating the stress tensor over its surface, the instantaneous total force and torque increase by as well, which are still zero.

Thus, scaling both the swimmer's motion and the motion of the surrounding fluid scales by the same factor, we still obtain a motion that respects the Stokes equation.

Proof by vector calculus

The proof of the scallop theorem can be represented in a mathematically elegant way. To do this, we must first understand the mathematical consequences of the linearity of Stokes equations. To summarize, the linearity of Stokes equations allows us to use the reciprocal theorem to relate the swimming velocity of the swimmer to the velocity field of the fluid around its surface (known as the swimming gait), which changes according to the periodic motion it exhibits. This relation allows us to conclude that locomotion is independent of swimming rate. Subsequently, this leads to the discovery that reversal of periodic motion is identical to the forward motion due to symmetry, allowing us to conclude that there can be no net displacement. [2]

Rate-independence

The reciprocal theorem describes the relationship between two Stokes flows in the same geometry where inertial effects are insignificant compared to viscous effects. Consider a fluid filled region bounded by surface with a unit normal . Suppose we have solutions to Stokes equations in the domain possessing the form of the velocity fields and . The velocity fields harbor corresponding stress fields and respectively. Then the following equality holds:

The reciprocal theorem allows us to obtain information about a certain flow by using information from another flow. This is preferable to solving Stokes equations, which is difficult due to not having a known boundary condition. This particularly useful if one wants to understand flow from a complicated problem by studying the flow of a simpler problem in the same geometry.

One can use the reciprocal theorem to relate the swimming velocity, , of a swimmer subject to a force to its swimming gait :

Now that we have established that the relationship between the instantaneous swimming velocity in the direction of the force acting on the body and its swimming gait follow the general form

where and denote the positions of points on the surface of the swimmer, we can establish that locomotion is independent of rate. Consider a swimmer that deforms in a periodic fashion through a sequence of motions between the times and The net displacement of the swimmer is

Now consider the swimmer deforming in the same manner but at a different rate. We describe this with the mapping

Using this mapping, we see that

This result means that the net distance traveled by the swimmer does not depend on the rate at which it is being deformed, but only on the geometrical sequence of shape. This is the first key result.

Symmetry of forward and backward motion

If a swimmer is moving in a periodic fashion that is time invariant, we know that the average displacement during one period must be zero. To illustrate the proof, let us consider a swimmer deforming during one period that starts and ends at times and . That means its shape at the start and end are the same, i.e. Next, we consider motion obtained by time-reversal symmetry of the first motion that occurs during the period starting and ending at times and using a similar mapping as in the previous section, we define and and define the shape in the reverse motion to be the same as the shape in the forward motion, Now we find the relationship between the net displacements in these two cases:

This is the second key result. Combining with our first key result from the previous section, we see that We see that a swimmer that reverses its motion by reversing its sequence of shape changes leads to the opposite distance traveled. In addition, since the swimmer exhibits reciprocal body deformation, the sequence of motion is the same between and and and Thus, the distance traveled should be the same independently of the direction of time, meaning that reciprocal motion cannot be used for net motion in low Reynolds number environments.

Exceptions

The scallop theorem holds if we assume that a swimmer undergoes reciprocal motion in an infinite quiescent Newtonian fluid in the absence of inertia and external body forces. However, there are instances where the assumptions for the scallop theorem are violated. [3] In one case, successful swimmers in viscous environments must display non-reciprocal body kinematics. In another case, if a swimmer is in a non-Newtonian fluid, locomotion can be achieved as well.

Types of non-reciprocal motion

In his original paper, Purcell proposed a simple example of non-reciprocal body deformation, now commonly known as the Purcell swimmer. This simple swimmer possess two degrees of freedom for motion: a two-hinged body composed of three rigid links rotating out-of-phase with each other. However, any body with more than one degree of freedom of motion can achieve locomotion as well.

In general, microscopic organisms like bacteria have evolved different mechanisms to perform non-reciprocal motion:

Geometrically, the rotating flagellum is a one-dimensional swimmer, and it works because its motion is going around a circle-shaped configuration space, and a circle is not a reciprocating motion. The flexible arm is a multi-dimensional swimmer, and it works because its motion is going around a circle in a square-shaped configuration space. Notice that the first kind of motion has nontrivial homotopy, but the second kind has trivial homotopy.

Non-Newtonian fluids

The assumption of a Newtonian fluid is essential since Stokes equations will not remain linear and time-independent in an environment that possesses complex mechanical and rheological properties. It is also common knowledge that many living microorganisms live in complex non-Newtonian fluids, which are common in biologically relevant environments. For instance, crawling cells often migrate in elastic polymeric fluids. Non-Newtonian fluids have several properties that can be manipulated to produce small scale locomotion. [3]

First, one such exploitable property is normal stress differences. These differences will arise from the stretching of the fluid by the flow of the swimmer. Another exploitable property is stress relaxation. Such time evolution of such stresses contain a memory term, though the extent in which this can be utilized is largely unexplored. Last, non-Newtonian fluids possess viscosities that are dependent on the shear rate. In other words, a swimmer would experience a different Reynolds number environment by altering its rate of motion. Many biologically relevant fluids exhibit shear-thinning, meaning viscosity decreases with shear rate. In such an environment, the rate at which a swimmer exhibits reciprocal motion would be significant as it would no longer be time invariant. This is in stark contrast to what we established where the rate in which a swimmer moves is irrelevant for establishing locomotion. Thus, a reciprocal swimmer can be designed in a non-Newtonian fluid. Qiu et al. (2014) were able to design a micro scallop in a non-Newtonian fluid. [7]

See also

Related Research Articles

<span class="mw-page-title-main">Fluid dynamics</span> Aspects of fluid mechanics involving flow

In physics, physical chemistry and engineering, fluid dynamics is a subdiscipline of fluid mechanics that describes the flow of fluids—liquids and gases. It has several subdisciplines, including aerodynamics and hydrodynamics. Fluid dynamics has a wide range of applications, including calculating forces and moments on aircraft, determining the mass flow rate of petroleum through pipelines, predicting weather patterns, understanding nebulae in interstellar space and modelling fission weapon detonation.

<span class="mw-page-title-main">Lorentz force</span> Force acting on charged particles in electric and magnetic fields

In physics, specifically in electromagnetism, the Lorentz force is the combination of electric and magnetic force on a point charge due to electromagnetic fields. A particle of charge q moving with a velocity v in an electric field E and a magnetic field B experiences a force of

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances. They were named after French engineer and physicist Claude-Louis Navier and the Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842–1850 (Stokes).

In vector calculus, the divergence theorem, also known as Gauss's theorem or Ostrogradsky's theorem, is a theorem relating the flux of a vector field through a closed surface to the divergence of the field in the volume enclosed.

<span class="mw-page-title-main">Stream function</span> Function for incompressible divergence-free flows in two dimensions

In fluid dynamics, two types of stream function are defined:

A Newtonian fluid is a fluid in which the viscous stresses arising from its flow are at every point linearly correlated to the local strain rate — the rate of change of its deformation over time. Stresses are proportional to the rate of change of the fluid's velocity vector.

<span class="mw-page-title-main">Scalar potential</span> When potential energy difference depends only on displacement

In mathematical physics, scalar potential, simply stated, describes the situation where the difference in the potential energies of an object in two different positions depends only on the positions, not upon the path taken by the object in traveling from one position to the other. It is a scalar field in three-space: a directionless value (scalar) that depends only on its location. A familiar example is potential energy due to gravity.

<span class="mw-page-title-main">Onsager reciprocal relations</span> Relations between flows and forces, or gradients, in thermodynamic systems

In thermodynamics, the Onsager reciprocal relations express the equality of certain ratios between flows and forces in thermodynamic systems out of equilibrium, but where a notion of local equilibrium exists.

In physics and mathematics, the Helmholtz decomposition theorem or the fundamental theorem of vector calculus states that any sufficiently smooth, rapidly decaying vector field in three dimensions can be resolved into the sum of an irrotational (curl-free) vector field and a solenoidal (divergence-free) vector field. This is named after Hermann von Helmholtz.

<span class="mw-page-title-main">Stokes flow</span> Type of fluid flow

Stokes flow, also named creeping flow or creeping motion, is a type of fluid flow where advective inertial forces are small compared with viscous forces. The Reynolds number is low, i.e. . This is a typical situation in flows where the fluid velocities are very slow, the viscosities are very large, or the length-scales of the flow are very small. Creeping flow was first studied to understand lubrication. In nature, this type of flow occurs in the swimming of microorganisms and sperm. In technology, it occurs in paint, MEMS devices, and in the flow of viscous polymers generally.

Fluid mechanics is the branch of physics concerned with the mechanics of fluids and the forces on them. It has applications in a wide range of disciplines, including mechanical, aerospace, civil, chemical, and biomedical engineering, as well as geophysics, oceanography, meteorology, astrophysics, and biology.

In numerical methods, total variation diminishing (TVD) is a property of certain discretization schemes used to solve hyperbolic partial differential equations. The most notable application of this method is in computational fluid dynamics. The concept of TVD was introduced by Ami Harten.

In fluid mechanics, potential vorticity (PV) is a quantity which is proportional to the dot product of vorticity and stratification. This quantity, following a parcel of air or water, can only be changed by diabatic or frictional processes. It is a useful concept for understanding the generation of vorticity in cyclogenesis, especially along the polar front, and in analyzing flow in the ocean.

In fluid mechanics, Kelvin's circulation theorem states:

In a barotropic, ideal fluid with conservative body forces, the circulation around a closed curve moving with the fluid remains constant with time.

The derivation of the Navier–Stokes equations as well as its application and formulation for different families of fluids, is an important exercise in fluid dynamics with applications in mechanical engineering, physics, chemistry, heat transfer, and electrical engineering. A proof explaining the properties and bounds of the equations, such as Navier–Stokes existence and smoothness, is one of the important unsolved problems in mathematics.

The Cauchy momentum equation is a vector partial differential equation put forth by Cauchy that describes the non-relativistic momentum transport in any continuum.

<span class="mw-page-title-main">Reynolds number</span> Ratio of inertial to viscous forces acting on a liquid

In fluid dynamics, the Reynolds number is a dimensionless quantity that helps predict fluid flow patterns in different situations by measuring the ratio between inertial and viscous forces. At low Reynolds numbers, flows tend to be dominated by laminar (sheet-like) flow, while at high Reynolds numbers, flows tend to be turbulent. The turbulence results from differences in the fluid's speed and direction, which may sometimes intersect or even move counter to the overall direction of the flow. These eddy currents begin to churn the flow, using up energy in the process, which for liquids increases the chances of cavitation.

<span class="mw-page-title-main">Stokes' theorem</span> Theorem in vector calculus

Stokes' theorem, also known as the Kelvin–Stokes theorem after Lord Kelvin and George Stokes, the fundamental theorem for curls or simply the curl theorem, is a theorem in vector calculus on . Given a vector field, the theorem relates the integral of the curl of the vector field over some surface, to the line integral of the vector field around the boundary of the surface. The classical theorem of Stokes can be stated in one sentence: The line integral of a vector field over a loop is equal to the surface integral of its curl over the enclosed surface. It is illustrated in the figure, where the direction of positive circulation of the bounding contour ∂Σ, and the direction n of positive flux through the surface Σ, are related by a right-hand-rule. For the right hand the fingers circulate along ∂Σ and the thumb is directed along n.

The viscous vortex domains (VVD) method is a mesh-free method of computational fluid dynamics for directly numerically solving 2D Navier-Stokes equations in Lagrange coordinates. It doesn't implement any turbulence model and free of arbitrary parameters. The main idea of this method is to present vorticity field with discrete regions (domains), which travel with diffusive velocity relatively to fluid and conserve their circulation. The same approach was used in Diffusion Velocity method of Ogami and Akamatsu, but VVD uses other discrete formulas

In fluid dynamics, Prandtl–Batchelor theorem states that if in a two-dimensional laminar flow at high Reynolds number closed streamlines occur, then the vorticity in the closed streamline region must be a constant. A similar statement holds true for axisymmetric flows. The theorem is named after Ludwig Prandtl and George Batchelor. Prandtl in his celebrated 1904 paper stated this theorem in arguments, George Batchelor unaware of this work proved the theorem in 1956. The problem was also studied in the same year by Richard Feynman and Paco Lagerstrom and by W.W. Wood in 1957.

References

  1. 1 2 Purcell, E. M. (1977), "Life at low reynolds number", American Journal of Physics, 45 (1): 3–11, Bibcode:1977AmJPh..45....3P, doi:10.1119/1.10903, hdl: 2433/226838
  2. 1 2 Lauga, Eric; Powers, Thomas R. (2009), "The hydrodynamics of swimming microorganisms", Reports on Progress in Physics, 72 (9): 096601, arXiv: 0812.2887 , Bibcode:2009RPPh...72i6601L, doi:10.1088/0034-4885/72/9/096601, S2CID   3932471
  3. 1 2 Lauga, Eric (2011), "Life around the scallop theorem", Soft Matter, 7 (7): 3060–3065, arXiv: 1011.3051 , Bibcode:2011SMat....7.3060L, doi:10.1039/C0SM00953A, S2CID   96762619
  4. Berg HC & Anderson RA (1973). "Bacteria swim by rotating their flagellar filaments". Nature. 245 (5425): 380–382. Bibcode:1973Natur.245..380B. doi:10.1038/245380a0. PMID   4593496. S2CID   4173914.
  5. Silverman M & Simon M (1974). "Flagellar rotation and the mechanism of bacterial motility". Nature. 249 (100): 73–74. Bibcode:1974Natur.249...73S. doi:10.1038/249073a0. PMID   4598030. S2CID   10370084.
  6. Brokaw CJ (1991). "Microtubule Sliding in Swimming Sperm Flagella: Direct and Indirect Measurements on Sea Urchin and Tunicate Spermatozoa". J Cell Biol. 114 (6): 1201–1215. doi:10.1083/jcb.114.6.1201. PMC   2289132 . PMID   1894694.
  7. Qiu, Tian; Lee, Tung-Chun; Mark, Andrew G.; Morozov, Konstantin I.; Münster, Raphael; Mierka, Otto; Turek, Stefan; Leshansky, Alexander M.; Fischer, Peer (2014), "Swimming by reciprocal motion at low Reynolds number", Nature Communications, 5: 5119, Bibcode:2014NatCo...5.5119Q, doi: 10.1038/ncomms6119 , PMC   4241991 , PMID   25369018