Flotation of flexible objects

Last updated

Flotation of flexible objects is a phenomenon in which the bending of a flexible material allows an object to displace a greater amount of fluid than if it were completely rigid. This ability to displace more fluid translates directly into an ability to support greater loads, giving the flexible structure an advantage over a similarly rigid one. Inspiration to study the effects of elasticity are taken from nature, where plants, such as black pepper, and animals living at the water surface have evolved to take advantage of the load-bearing benefits elasticity imparts.

Contents

History

In his work "On Floating Bodies", Archimedes [1] famously stated:

Any object, wholly or partially immersed in a fluid, is buoyed up by a force equal to the weight of the fluid displaced by the object.

Archimedes of Syracuse

While this basic idea carried enormous weight and has come to form the basis of understanding why objects float, it is best applied for objects with a characteristic length scale greater than the capillary length. What Archimedes had failed to predict was the influence of surface tension and its impact at small length scales.

More recent works, such as that of Keller, [2] have extended these principles by considering the role of surface tension forces on partially submerged bodies. Keller, for instance, demonstrated analytically that the weight of water displaced by a meniscus is equal to the vertical component of the surface tension force.

The well known phenomena by which water striders support themselves through surface tension effects can be re-examined by considering the compliance and resultant deformation of the insect's leg at the water line. Water strider.jpg
The well known phenomena by which water striders support themselves through surface tension effects can be re-examined by considering the compliance and resultant deformation of the insect's leg at the water line.

Nonetheless, the role of flexibility and its impact on an object's load-bearing potential is one that did receive attention until the mid-2000s and onward. In an initial study, Vella [3] studied the load supported by a raft composed of thin, rigid strips. Specifically, he compared the case of floating individual strips to floating an aggregation of strips, wherein the aggregate structure causes portions of the meniscus (and hence, resulting surface tension force) to disappear. By extending his analysis to consider a similar system composed of thin strips of some finite bending stiffness, he found that this later case in fact was able support a greater load. [4]

A well known work in the area of surface tension aided flotation was the analysis of water strider locomotion along the surface of water. [5] Using the idea of flexible structures, [6] Ji et al. re-examined this problem by considering the compliance of a water strider leg. By modeling the leg as a compliant structure that deforms at the water surface (rather than pierce it), Ji was able to ascertain what added benefit this flexibility has in supporting the insect. Other studies on the water strider have examined the ways in which flexibility can affect wetting properties of the leg. [7]

Another track of research has been to investigate how exactly the interaction between liquid and a compliant object leads to the resulting deformation. In one example, such analysis has been extended to explain the difficulty in submerging hairs in a fluid. [8] These works focus on behavior near the contact line, and consider what role non-linear effects such as slippage play.

Physical explanation of phenomena

In a liquid solution, any given liquid molecule experience strong cohesive forces from neighboring molecules. While these forces are balanced in the bulk, molecules at the surface of the solution are surrounded on one side by water molecules and on the other side by gas molecules. The resulting imbalance of cohesive forces along the surface results in a net "pull" toward the bulk, giving rise to the phenomena of surface tension.

Rigid plate submerged below water line. If the plate is flexible, its deformation (dashed) results in the displacement of additional fluid (blue, diagonal). However, doing so results in a simultaneous loss due to the narrowing of the above column (red, vertical). Flexible plate displaced below water line.jpg
Rigid plate submerged below water line. If the plate is flexible, its deformation (dashed) results in the displacement of additional fluid (blue, diagonal). However, doing so results in a simultaneous loss due to the narrowing of the above column (red, vertical).

When a hydrophobic object of weight is placed on the surface of water, its weight begins deforming the water line. The hydrophobic nature of the object means that the water will attempt to minimize contact due to an unfavorable energy tradeoff associated with wetting. As a result, surface tension attempts to pull back on the water line in order to minimize contact with the hydrophobic object and retain a lowest energy state. This action by the surface to pull back on the depressed water interface is the source of a capillary force, which acts tangentially along the contact line and thereby gives rise to a component in the vertical direction. An attempt to further depress the object is resisted by this capillary force until the contact line reaches a vertical position located about two capillary lengths below the undisturbed water line. [9] Once this occurs, the meniscus collapses and the object sinks.

The more fluid a floating object is able to displace, the greater the load it is able to bear. As a result, the ultimate payoff of flexibility is in determining whether or not a bent configuration results in an increased volume of displaced water. As a flexible object bends, it penetrates further into the water and increases the total fluid displaced above it. However, this bending action necessarily forces the cross-section at the water line to decrease, narrowing the column of displaced water above the object. Thus, whether or not bending is advantageous is ultimately given by a tradeoff of these factors.

Mathematical model: flotation of hinged plates

Two rigid plates connected by a torsional spring

The following analysis is taken largely from the work of Burton and Bush, [9] and offers some mathematical insight into the role that flexibility plays in improving load-bearing characteristics of floating objects.

Consider two plates of infinite width, thickness , and length that are connected by a torsional spring with spring constant per unit width . Furthermore, let be the angle between a plate and the horizontal, and the from where the meniscus meets the plate to the horizontal. The distance from the undisturbed water line to the plate's outer edge is . The density of water is , the density of air is considered negligible, and the plate density, , shall be varied. All systems naturally assume a configuration that minimizes total energy. Thus, the goal of this analysis is to identify the configurations (i.e., values of and ) that result in a stable equilibrium for a given value of .

For a total system energy of , it is natural to distinguish sub-components:

Two rigid plates connected by a torsional spring and submerged between the water line. Since the material is taken to be hydrophobic, the water deforms, forming the meniscus at the plate edge. Submering hinged plates.jpg
Two rigid plates connected by a torsional spring and submerged between the water line. Since the material is taken to be hydrophobic, the water deforms, forming the meniscus at the plate edge.


: Work done on the system
: System potential energy

In defining , there are several associated components:


is the work done on the interface by hydrostatic pressure
is the work done on the plates by hydrostatic pressure
is the work done on the plates by gravitational force
is the work done on the plates by surface tension forces

Similarly, the system potential energy, , is taken to be composed of two terms:

is the surface energy of the water/air interface
is the energy stored in the torsional spring and is equal to

There are two ways in which the system energy can change by an incremental amount. The first is a translation of the center of mass of the plates by some distance . The second is an incremental change, in the hinge angle. Such a change will induce a new moment.

As mentioned, the system will seek the orientation that minimizes in order to find point of stable equilibrium. Writing out these terms more explicitly:


Here, is the equation air/water interface, is the incremental displacement of the interface, and is the surface tension of water.

For a given value of , stable equilibrium configurations are identified as being those values of and that satisfy



Taken in a different light, these conditions can be seen as identifying and that result in zero net force and zero net torque for a given .

Analytical results for maximum load

Defining non-dimensional plate length , non-dimensional plate edge depth , and non-dimensional load , Burton and Bush derived the following analytical results:



The equations for and give the configuration parameters that give the maximum value of . For further insight, it is helpful to examine various regimes of the non-dimensional plate length, .

Case 1: Small scale β << H

When the characteristic plate length is much smaller than the characteristic plate edge depth, the effects of gravity, surface tension, and spring energy become dominant. In this limiting case, it turns out that flexibility does not improve load-bearing capabilities of the plates; indeed, the optimal configuration is a flat plate. Since the plate length is so much smaller than the displacement from the undisturbed water line, the extra fluid displaced by bending a rigid plate is outweighed by the loss of fluid in the column above the plate.

Case 2: Mid scale β ~ H

In this regime, flexibility may or may not improve load-bearing capabilities of the plates. The two characteristics lengths are of comparable dimension, so particular values for each determine whether or not additional fluid displaced through bending exceeds fluid lost through the narrowing of the column.

Case 3: Large scale β >> H

In this regime, the benefit of flexibility is most pronounced. The characteristic plate length is significantly longer than the characteristic depth to which the plate is submerged beneath the water line. As a result, the narrowing column above the plate is negligible, which the additional displacement of water due to bending is significant.

Extensions to continuously deformable body

To relate this mathematical to physical systems, the above analysis can be extended to continuously deformable bodies. Generalizing the equations of the two plate system allows one to write a corresponding set of equations for the case of a continuously deformable plate. This continuously deformable plate is composed of sub-plates, where similar force and torque equilibrium conditions described before must be satisfied for each sub-plate. Such analysis reveals that for a highly compliant 2D structure with a characteristic length much greater than the capillary length, the shape bearing the highest load is a perfect semi-circle. As stiffness increases, the semi-circle is deformed to shapes with lower curvature.

This initial look at continuously deformable bodies represents an initial stab into a very complicated problem. With the groundwork laid here in this analysis, it is likely that future works will implement this general ideology in a finite element approach. Doing so will allow much closer simulation of real world phenomena and will aid in determining how effects of elasticity can aid in the design of robots, instruments, and other devices that operate along the water line.

Examples in nature

Ants

To survive floods, colonies of fire ants will join limbs to create a living raft. Fire ants cluster in water.jpg
To survive floods, colonies of fire ants will join limbs to create a living raft.

Several species of ants including fire ants [10] [11] , formica [12] and red imported fire ant [13] [14] , have developed adaptation to survive floods that threaten the colony. While individual fire ants are hydrophobic and flounder at the waters surface, large groups of ants can link together to form a temporary living clump, known as raft, and float along the waterline for as many as 12 days until reaching dry land. [14] These clumps can include as many as 100,000 individual ants while the queen and larvae are evacuated from the flooding colony sitting upon this living raft. [10] Ants clumped in this way will recognize different fluid flow conditions and adapt their behavior accordingly to preserve the raft's stability. [11]

The importance of flexibility in this self-assembled ant raft is several fold. The extra weight-bearing that flexibility imparts is vital as hungry fish will swim along the underside of the raft and eat at many of the members. Furthermore, as waves travel along the water surface, the ant raft's flexibility allows it to effectively "roll" with the wave and minimize disturbances it would otherwise cause for a similar but rigid structure.

Aquatic vegetation

To protect its reproductive material, the daisy's petals fold in during floods to prevent the rush of water. Dscn3156-daisy-water 1200x900.jpg
To protect its reproductive material, the daisy's petals fold in during floods to prevent the rush of water.

Among aquatic vegetation, the lily pad is perhaps the most recognizable, commonly associated with ponds and lakes. Their flexibility allows for increased loads, enabling them to support animals, such as frogs, many times their own weight.

Some aquatic flowers, such as the daisy Bellis perennis, use compliance as a survival mechanism. Such flowers have roots that extend down to the underlying soil, anchoring the flower to the surface of the water. When flooding occurs, the petals pull inward and deform the water line, shielding the genetic material in the core. [15] Some flowers are even known to completely close up into a shell in this fashion, trapping air inside.

Related Research Articles

In fluid mechanics, the Grashof number is a dimensionless number which approximates the ratio of the buoyancy to viscous forces acting on a fluid. It frequently arises in the study of situations involving natural convection and is analogous to the Reynolds number.

In fluid mechanics, the Rayleigh number (Ra, after Lord Rayleigh) for a fluid is a dimensionless number associated with buoyancy-driven flow, also known as free (or natural) convection. It characterises the fluid's flow regime: a value in a certain lower range denotes laminar flow; a value in a higher range, turbulent flow. Below a certain critical value, there is no fluid motion and heat transfer is by conduction rather than convection. For most engineering purposes, the Rayleigh number is large, somewhere around 106 to 108.

<span class="mw-page-title-main">Stress–energy tensor</span> Tensor describing energy momentum density in spacetime

The stress–energy tensor, sometimes called the stress–energy–momentum tensor or the energy–momentum tensor, is a tensor physical quantity that describes the density and flux of energy and momentum in spacetime, generalizing the stress tensor of Newtonian physics. It is an attribute of matter, radiation, and non-gravitational force fields. This density and flux of energy and momentum are the sources of the gravitational field in the Einstein field equations of general relativity, just as mass density is the source of such a field in Newtonian gravity.

In the mathematical field of differential geometry, the Riemann curvature tensor or Riemann–Christoffel tensor is the most common way used to express the curvature of Riemannian manifolds. It assigns a tensor to each point of a Riemannian manifold. It is a local invariant of Riemannian metrics which measures the failure of the second covariant derivatives to commute. A Riemannian manifold has zero curvature if and only if it is flat, i.e. locally isometric to the Euclidean space. The curvature tensor can also be defined for any pseudo-Riemannian manifold, or indeed any manifold equipped with an affine connection.

<span class="mw-page-title-main">Compressibility</span> Measure of the relative volume change of a fluid or solid as a response to a pressure change

In thermodynamics and fluid mechanics, the compressibility is a measure of the instantaneous relative volume change of a fluid or solid as a response to a pressure change. In its simple form, the compressibility may be expressed as

In the calculus of variations, a field of mathematical analysis, the functional derivative relates a change in a functional to a change in a function on which the functional depends.

<span class="mw-page-title-main">String vibration</span> A wave

A vibration in a string is a wave. Resonance causes a vibrating string to produce a sound with constant frequency, i.e. constant pitch. If the length or tension of the string is correctly adjusted, the sound produced is a musical tone. Vibrating strings are the basis of string instruments such as guitars, cellos, and pianos.

<span class="mw-page-title-main">Bending</span> Strain caused by an external load

In applied mechanics, bending characterizes the behavior of a slender structural element subjected to an external load applied perpendicularly to a longitudinal axis of the element.

In physics, precisely in the study of the theory of general relativity and many alternatives to it, the post-Newtonian formalism is a calculational tool that expresses Einstein's (nonlinear) equations of gravity in terms of the lowest-order deviations from Newton's law of universal gravitation. This allows approximations to Einstein's equations to be made in the case of weak fields. Higher-order terms can be added to increase accuracy, but for strong fields, it may be preferable to solve the complete equations numerically. Some of these post-Newtonian approximations are expansions in a small parameter, which is the ratio of the velocity of the matter forming the gravitational field to the speed of light, which in this case is better called the speed of gravity. In the limit, when the fundamental speed of gravity becomes infinite, the post-Newtonian expansion reduces to Newton's law of gravity.

A quasiprobability distribution is a mathematical object similar to a probability distribution but which relaxes some of Kolmogorov's axioms of probability theory. Quasiprobabilities share several of general features with ordinary probabilities, such as, crucially, the ability to yield expectation values with respect to the weights of the distribution. However, they can violate the σ-additivity axiom: integrating over them does not necessarily yield probabilities of mutually exclusive states. Indeed, quasiprobability distributions also have regions of negative probability density, counterintuitively, contradicting the first axiom. Quasiprobability distributions arise naturally in the study of quantum mechanics when treated in phase space formulation, commonly used in quantum optics, time-frequency analysis, and elsewhere.

<span class="mw-page-title-main">Covariant formulation of classical electromagnetism</span> Ways of writing certain laws of physics

The covariant formulation of classical electromagnetism refers to ways of writing the laws of classical electromagnetism in a form that is manifestly invariant under Lorentz transformations, in the formalism of special relativity using rectilinear inertial coordinate systems. These expressions both make it simple to prove that the laws of classical electromagnetism take the same form in any inertial coordinate system, and also provide a way to translate the fields and forces from one frame to another. However, this is not as general as Maxwell's equations in curved spacetime or non-rectilinear coordinate systems.

<span class="mw-page-title-main">Maxwell's equations in curved spacetime</span> Electromagnetism in general relativity

In physics, Maxwell's equations in curved spacetime govern the dynamics of the electromagnetic field in curved spacetime or where one uses an arbitrary coordinate system. These equations can be viewed as a generalization of the vacuum Maxwell's equations which are normally formulated in the local coordinates of flat spacetime. But because general relativity dictates that the presence of electromagnetic fields induce curvature in spacetime, Maxwell's equations in flat spacetime should be viewed as a convenient approximation.

The harmonic coordinate condition is one of several coordinate conditions in general relativity, which make it possible to solve the Einstein field equations. A coordinate system is said to satisfy the harmonic coordinate condition if each of the coordinate functions xα satisfies d'Alembert's equation. The parallel notion of a harmonic coordinate system in Riemannian geometry is a coordinate system whose coordinate functions satisfy Laplace's equation. Since d'Alembert's equation is the generalization of Laplace's equation to space-time, its solutions are also called "harmonic".

In many-body theory, the term Green's function is sometimes used interchangeably with correlation function, but refers specifically to correlators of field operators or creation and annihilation operators.

In mechanics and geodynamics, a critical taper is the equilibrium angle made by the far end of a wedge-shaped agglomeration of material that is being pushed by the near end. The angle of the critical taper is a function of the material properties within the wedge, pore fluid pressure, and strength of the fault along the base of the wedge.

In continuum mechanics, a compatible deformation tensor field in a body is that unique tensor field that is obtained when the body is subjected to a continuous, single-valued, displacement field. Compatibility is the study of the conditions under which such a displacement field can be guaranteed. Compatibility conditions are particular cases of integrability conditions and were first derived for linear elasticity by Barré de Saint-Venant in 1864 and proved rigorously by Beltrami in 1886.

Paleosalinity is the salinity of the global ocean or of an ocean basin at a point in geological history.

<span class="mw-page-title-main">Kirchhoff–Love plate theory</span>

The Kirchhoff–Love theory of plates is a two-dimensional mathematical model that is used to determine the stresses and deformations in thin plates subjected to forces and moments. This theory is an extension of Euler-Bernoulli beam theory and was developed in 1888 by Love using assumptions proposed by Kirchhoff. The theory assumes that a mid-surface plane can be used to represent a three-dimensional plate in two-dimensional form.

<span class="mw-page-title-main">Dynamic fluid film equations</span>

Fluid films, such as soap films, are commonly encountered in everyday experience. A soap film can be formed by dipping a closed contour wire into a soapy solution as in the figure on the right. Alternatively, a catenoid can be formed by dipping two rings in the soapy solution and subsequently separating them while maintaining the coaxial configuration.

In the mathematical field of differential geometry, a biharmonic map is a map between Riemannian or pseudo-Riemannian manifolds which satisfies a certain fourth-order partial differential equation. A biharmonic submanifold refers to an embedding or immersion into a Riemannian or pseudo-Riemannian manifold which is a biharmonic map when the domain is equipped with its induced metric. The problem of understanding biharmonic maps was posed by James Eells and Luc Lemaire in 1983. The study of harmonic maps, of which the study of biharmonic maps is an outgrowth, had been an active field of study for the previous twenty years. A simple case of biharmonic maps is given by biharmonic functions.

References

  1. Acott, Chris (1999). "The diving "Law-ers": A brief resume of their lives". South Pacific Underwater Medicine Society Journal. 29 (1). ISSN   0813-1988. OCLC   16986801. Archived from the original on 2011-04-02. Retrieved 2013-12-12.{{cite journal}}: CS1 maint: unfit URL (link)
  2. Keller, Joseph B. (1998). "Surface tension force on a partly submerged body". Physics of Fluids. 10 (11): 3009. Bibcode:1998PhFl...10.3009K. doi:10.1063/1.869820.
  3. Vella, Dominic; Metcalfe, Paul D.; Whittaker, Robert J. (2006). "Equilibrium Conditions for the Floating of Multiple Interfacial Objects". Journal of Fluid Mechanics. 549 (215): 215–224. arXiv: physics/0509009 . Bibcode:2006JFM...549..215V. doi:10.1017/s0022112005008013. S2CID   16347764.
  4. Vella, Dominic (2007). The fluid mechanics of floating and sinking (PhD Dissertation). Cambridge: Trinity College, University of Cambridge.
  5. Hu, David; Chan, Brian; Bush, John W.M. (2003). "The hydrodynamics of water strider locomotion". Nature. 424 (6949): 663–666. Bibcode:2003Natur.424..663H. doi:10.1038/nature01793. PMID   12904790. S2CID   4362791.
  6. Ji, Xiang-Ying; Wang, Jia-Wen; Feng, Xi-Xiao (2012). "Role of flexibility in the water repellency of water strider legs: Theory and experiment". Physical Review E. 85 (2): 021607. Bibcode:2012PhRvE..85b1607J. doi:10.1103/physreve.85.021607. PMID   22463223.
  7. Park, Kun Joong; Kim, Ho-Young (2008). "Bending of floating flexible legs". Journal of Fluid Mechanics. 610: 381–390. Bibcode:2008JFM...610..381P. doi:10.1017/s0022112008002784. S2CID   43925050.
  8. Anderotti, B.; Marchand, A.; Das, S.; Snoeijer, J.H. (2011). "Elastocapillary instability under partial wetting conditions: Bending versus buckling". Physical Review E. 84 (6): 061601. arXiv: 1109.3414 . Bibcode:2011PhRvE..84f1601A. doi:10.1103/physreve.84.061601. PMID   22304096. S2CID   27689511.
  9. 1 2 Burton, L.J.; Bush, J.W.M (2012). "Can flexibility help you float?" (PDF). Physics of Fluids. 24 (10): 101701–101701–7. Bibcode:2012PhFl...24j1701B. doi:10.1063/1.4757121. hdl: 1721.1/80721 . S2CID   17813611.
  10. 1 2 "Floating fire ants form rafts in Houston floodwaters". BBC. London. August 2017. Retrieved 30 August 2017.
  11. 1 2 Ouellette, Jennifer (16 September 2022). "Fire ant rafts form because of the Cheerios effect, study concludes". Ars Technica. Retrieved 18 October 2022.
  12. Avril, Amaury; Purcell, Jessica; Chapuisat, Michel (2016-04-07). "Ant workers exhibit specialization and memory during raft formation" (PDF). The Science of Nature. 103 (5–6): 36. Bibcode:2016SciNa.103...36A. doi:10.1007/s00114-016-1360-5. ISSN   0028-1042. PMID   27056046. S2CID   17142619.
  13. Mlot, N.J.; Craig, A.T.; Hu, D.L. (2011). "Fire ants self-assemble into waterproof rafts to survive floods". Proceedings of the National Academy of Sciences. 108 (19): 7669–7673. Bibcode:2011PNAS..108.7669M. doi: 10.1073/pnas.1016658108 . PMC   3093451 . PMID   21518911.
  14. 1 2 Adams, B.J.; Hooper-Bùi, L.M.; Strecker, R.M.; O'Brien, D.M. (2011). "Raft formation by the red imported fire ant, Solenopsis invicta". Journal of Insect Science. 11 (171): 171. doi:10.1673/031.011.17101. PMC   3462402 . PMID   22950473.
  15. Reis, Pedro M.; Hure, Jérémy; Jung, Sungwan; Bush, John W.M.; Clanet, Cristophe (2010). "Grabbing water". Soft Matter. 6 (22): 5705–5708. arXiv: 1207.3756 . Bibcode:2010SMat....6.5705R. doi:10.1039/c0sm00895h. hdl:1721.1/77954. S2CID   10173213.