Vous êtes sur la page 1sur 33

Newton's laws of motion

From Wikipedia, the free encyclopedia


Jump to: navigation, search For other uses, see Laws of motion. Newton's First and Second laws, in Latin, from the original 1687 edition of the Principia Mathematica.
Classical mechanics

Newton's Second Law

History of ...

[show]Fundamental concepts [hide]Formulations Newtonian mechanics Lagrangian mechanics Hamiltonian mechanics [show]Branches [show]Scientists
This box: view talk edit

Newton's laws of motion are three physical laws that form the basis for classical mechanics. They are: 1. In the absence of force, a body either is at rest or moves in a straight line with constant speed. 2. A body experiencing a force F experiences an acceleration a related to F by F = ma, where m is the mass of the body. Alternatively, force is equal to the time derivative of momentum. 3. Whenever a first body exerts a force F on a second body, the second body exerts a force F on the first body. F and F are equal in magnitude and opposite in direction.[note 1] These laws describe the relationship between the forces acting on a body and the motion of that body. They were first compiled by Sir Isaac Newton in his work Philosophi Naturalis Principia Mathematica, first published on July 5, 1687.[1] Newton used them to explain and investigate the motion of many physical objects and systems.[2] For example, in the third volume of the text, Newton showed that these laws of motion, combined with his law of universal gravitation, explained Kepler's laws of planetary motion. First law There exists a set of inertial reference frames relative to which all particles with no net force acting on them will move without change in their velocity. This law is often simplified as "A body persists in a state of rest or of uniform motion unless acted upon by an external force." Newton's first law is often referred to as the law of inertia. Second law Observed from an inertial reference frame, the net force on a particle is equal to the time rate of change of its linear momentum: F = d(mv)/dt. Since by definition the mass of a particle is constant, this law is often stated as, "Force equals mass times acceleration (F = ma): the net force on an object is equal to the mass of the object multiplied by its acceleration." Third law Whenever a particle A exerts a force on another particle B, B simultaneously exerts a force on A with the same magnitude in the opposite direction. The strong form of the law further postulates that these two forces act along the same line. This law is

often simplified into the sentence, "To every action there is an equal and opposite reaction." In the given interpretation mass, acceleration, momentum, and (most importantly) force are assumed to be externally defined quantities. This is the most common, but not the only interpretation: one can consider the laws to be a definition of these quantities. Some authors interpret the first law as defining what an inertial reference frame is; from this point of view, the second law only holds when the observation is made from an inertial reference frame, and therefore the first law cannot be proved as a special case of the second. Other authors do treat the first law as a corollary of the second.[3] The explicit concept of an inertial frame of reference was not developed until long after Newton's death. At speeds approaching the speed of light the effects of special relativity must be taken into account.[note 2]

Newton's first law


Lex I: Corpus omne perseverare in statu suo quiescendi vel movendi uniformiter in directum, nisi quatenus a viribus impressis cogitur statum illum mutare. Every body persists in its state of being at rest or of moving uniformly straight forward, except insofar as it is compelled to change its state by force impressed.[4] Newton's first law is also called the law of inertia. It states that if the vector sum of all forces (that is, the net force) acting on an object is zero, then the acceleration of the object is zero and its velocity is constant. Consequently:

An object that is not moving will not move until a force acts upon it. An object that is moving will not change its velocity until a net force acts upon it.

The first point needs no comment, but the second seems to violate everyday experience. For example, a hockey puck sliding along ice does not move forever; rather, it slows and eventually comes to a stop. According to Newton's first law, the puck comes to a stop because of a net external force applied in the direction opposite to its motion. This net external force is due to a frictional force between the puck and the ice, as well as a frictional force between the puck and the air. If the ice were frictionless and the puck were traveling in a vacuum, the net external force on the puck would be zero and it would travel with constant velocity so long as its path were unobstructed. Implicit in the discussion of Newton's first law is the concept of an inertial reference frame, which for the purposes of Newtonian mechanics is defined to be a reference frame in which Newton's first law holds true. There is a class of frames of reference (called inertial frames) relative to which the motion of a particle not subject to forces is a straight line.[5] Newton placed the law of inertia first to establish frames of reference for which the other laws are applicable.[5][6] To understand why the laws are restricted to inertial frames, consider a ball at rest inside an airplane on a runway. From the perspective of an observer within the airplane (that is, from the airplane's frame of reference) the ball will appear to move backward as the plane accelerates forward. This motion appears to contradict Newton's second law (F = ma), since, from the point of view of the passengers, there appears to be no force acting on the ball that would cause it to move. However, Newton's first law does not apply: the stationary ball does not remain stationary in the absence of external force. Thus the reference frame of the airplane is not inertial, and Newton's second law does not hold in the form F = ma.[note 3]

History of the first law


Newton's first law is a restatement of what Galileo had already described and Newton gave credit to Galileo. It differs from Aristotle's view that all objects have a natural place in the universe. Aristotle believed that heavy objects like rocks wanted to be at rest on the Earth

and that light objects like smoke wanted to be at rest in the sky and the stars wanted to remain in the heavens. However, a key difference between Galileo's idea and Aristotle's is that Galileo realized that force acting on a body determines acceleration, not velocity. This insight leads to Newton's First Lawno force means no acceleration, and hence the body will maintain its velocity. The law of inertia apparently occurred to several different natural philosophers and scientists independently. The inertia of motion was described in the 3rd century BC by the Chinese philosopher Mo Tzu, and in the 11th century by the Muslim scientists Alhazen[7] and Avicenna.[8] The 17th century philosopher Ren Descartes also formulated the law, although he did not perform any experiments to confirm it. The first law was understood philosophically well before Newton's publication of the law.
[note 4]

Newton's second law


Newton's second law states that the force applied to a body produces a proportional acceleration; the relationship between the two is

where F is the force applied, m is the mass of the body, and a is the body's acceleration. If the body is subject to multiple forces at the same time, then the acceleration is proportional to the vector sum (that is, the net force):

The second law can also be shown to relate the net force and the momentum p of the body:

Therefore, Newton's second law also states that the net force is equal to the time derivative of the body's momentum:

Consistent with the first law, the time derivative of the momentum is non-zero when the momentum changes direction, even if there is no change in its magnitude (see time derivative). The relationship also implies the conservation of momentum: when the net force on the body is zero, the momentum of the body is constant. Both statements of the second law are valid only for constant-mass systems,[9][10][11] since any mass that is gained or lost by the system will cause a change in momentum that is not the result of an external force. A different equation is necessary for variable-mass systems. Newton's second law requires modification if the effects of special relativity are to be taken into account, since it is no longer true that momentum is the product of inertial mass and velocity.

Impulse
An impulse I occurs when a force F acts over an interval of time t, and it is given by[12][13]

Since force is the time derivative of momentum, it follows that

This relation between impulse and momentum is closer to Newton's wording of the second law.[14] Impulse is a concept frequently used in the analysis of collisions and impacts.[15]

Variable-mass systems
Variable-mass systems, like a rocket burning fuel and ejecting spent gases, are not closed and cannot be directly treated by making mass a function of time in the second law.[10] The reasoning, given in An Introduction to Mechanics by Kleppner and Kolenkow and other modern texts, is that Newton's second law applies fundamentally to particles.[11] In classical mechanics, particles by definition have constant mass. In case of a well-defined system of particles, Newton's law can be extended by summing over all the particles in the system:

where Fnet is the total external force on the system, M is the total mass of the system, and acm is the acceleration of the center of mass of the system. Variable-mass systems like a rocket or a leaking bucket cannot usually be treated as a system of particles, and thus Newton's second law cannot be applied directly. Instead, the general equation of motion for a body whose mass m varies with time by either ejecting or accreting mass is obtained by rearranging the second law and adding a term to account for the momentum carried by mass entering or leaving the system:[9]

where u is the relative velocity of the escaping or incoming mass with respect to the center of mass of the body. Under some conventions, the quantity udm/dt on the left-hand side is defined as a force (the force exerted on the body by the changing mass, such as rocket exhaust) and is included in the quantity F. Then, by substituting the definition of acceleration, the equation becomes

History of the second law


Newton's Latin wording for the second law is: Lex II: Mutationem motus proportionalem esse vi motrici impressae, et fieri secundum lineam rectam qua vis illa imprimitur. This was translated quite closely in Motte's 1729 translation as: LAW II: The alteration of motion is ever proportional to the motive force impress'd; and is made in the direction of the right line in which that force is impress'd. According to modern ideas of how Newton was using his terminology,[note 5] this is understood, in modern terms, as an equivalent of: The change of momentum of a body is proportional to the impulse impressed on the body, and happens along the straight line on which that impulse is impressed. Motte's 1729 translation of Newton's Latin continued with Newton's commentary on the second law of motion, reading: If a force generates a motion, a double force will generate double the motion, a triple force triple the motion, whether that force be impressed altogether and at once, or gradually and

successively. And this motion (being always directed the same way with the generating force), if the body moved before, is added to or subtracted from the former motion, according as they directly conspire with or are directly contrary to each other; or obliquely joined, when they are oblique, so as to produce a new motion compounded from the determination of both. The sense or senses in which Newton used his terminology, and how he understood the second law and intended it to be understood, have been extensively discussed by historians of science, along with the relations between Newton's formulation and modern formulations.[note 6]

Newton's third law: law of reciprocal actions

Newton's third law. The skaters' forces on each other are equal in magnitude, but act in opposite directions. Lex III: Actioni contrariam semper et qualem esse reactionem: sive corporum duorum actiones in se mutuo semper esse quales et in partes contrarias dirigi. ''To every action there is always an equal and opposite reaction: or the forces of two bodies on each other are always equal and are directed in opposite directions''. A more direct translation than the one just given above is: LAW III: To every action there is always opposed an equal reaction: or the mutual actions of two bodies upon each other are always equal, and directed to contrary parts. Whatever draws or presses another is as much drawn or pressed by that other. If you press a stone with your finger, the finger is also pressed by the stone. If a horse draws a stone tied to a rope, the horse (if I may so say) will be equally drawn back towards the stone: for the distended rope, by the same endeavour to relax or unbend itself, will draw the horse as much towards the stone, as it does the stone towards the horse, and will obstruct the progress of the one as much as it advances that of the other. If a body impinges upon another, and by its force changes the motion of the other, that body also (because of the equality of the mutual pressure) will undergo an equal change, in its own motion, toward the contrary part. The changes made by these actions are equal, not in the velocities but in the motions of the bodies; that is to say, if the bodies are not hindered by any other impediments. For, as the motions are equally changed, the changes of the velocities made toward contrary parts are reciprocally proportional to the bodies. This law takes place also in attractions, as will be proved in the next scholium.[note 7] In the above, as usual, motion is Newton's name for momentum, hence his careful distinction between motion and velocity. The Third Law means that all forces are interactions, and thus that there is no such thing as a unidirectional force. If body A exerts a force on body B, simultaneously, body B exerts a force of the same magnitude body A, both forces acting along the same line. As shown in the diagram opposite, the skaters' forces on each other are equal in magnitude, but act in opposite directions. Although the forces are equal, the accelerations are not: the less massive skater will have a greater acceleration due to Newton's second law. It is important to note that the action and reaction act on different objects and do not cancel each other out. The two forces in Newton's third law are of the same type (e.g., if the road exerts a forward frictional force on an accelerating car's tires, then it is also a frictional force that Newton's third law predicts for the tires pushing backward on the road). Newton used the third law to derive the law of conservation of momentum;[16] however from a deeper perspective, conservation of momentum is the more fundamental idea (derived via Noether's theorem from Galilean invariance), and holds in cases where Newton's third law appears to fail, for instance when force fields as well as particles carry momentum, and in quantum mechanics.

Importance and range of validity


Newton's laws were verified by experiment and observation for over 200 years, and they are excellent approximations at the scales and speeds of everyday life. Newton's laws of motion, together with his law of universal gravitation and the mathematical techniques of calculus, provided for the first time a unified quantitative explanation for a wide range of physical phenomena. These three laws hold to a good approximation for macroscopic objects under everyday conditions. However, Newton's laws (combined with Universal Gravitation and Classical Electrodynamics) are inappropriate for use in certain circumstances, most notably at very small scales, very high speeds (in special relativity, the Lorentz factor must be included in the expression for momentum along with rest mass and velocity) or very strong gravitational fields. Therefore, the laws cannot be used to explain phenomena such as conduction of electricity in a semiconductor, optical properties of substances, errors in nonrelativistically corrected GPS systems and superconductivity. Explanation of these phenomena requires more sophisticated physical theory, including General Relativity and Relativistic Quantum Mechanics. In quantum mechanics concepts such as force, momentum, and position are defined by linear operators that operate on the quantum state; at speeds that are much lower than the speed of light, Newton's laws are just as exact for these operators as they are for classical objects. At speeds comparable to the speed of light, the second law holds in the original form F = dp/dt, which says that the force is the derivative of the momentum of the object with respect to time, but some of the newer versions of the second law (such as the constant mass approximation above) do not hold at relativistic velocities.

Relationship to the conservation laws


In modern physics, the laws of conservation of momentum, energy, and angular momentum are of more general validity than Newton's laws, since they apply to both light and matter, and to both classical and non-classical physics. This can be stated simply, "Momentum, energy and angular momentum cannot be created or destroyed." Because force is the time derivative of momentum, the concept of force is redundant and subordinate to the conservation of momentum, and is not used in fundamental theories (e.g. quantum mechanics, quantum electrodynamics, general relativity, etc.). The standard model explains in detail how the three fundamental forces known as gauge forces originate out of exchange by virtual particles. Other forces such as gravity and fermionic degeneracy pressure also arise from the momentum conservation. Indeed, the conservation of 4momentum in inertial motion via curved space-time results in what we call gravitational force in general relativity theory. Application of space derivative (which is a momentum operator in quantum mechanics) to overlapping wave functions of pair of fermions (particles with semi-integer spin) results in shifts of maxima of compound wavefunction away from each other, which is observable as "repulsion" of fermions. Newton stated the third law within a world-view that assumed instantaneous action at a distance between material particles. However, he was prepared for philosophical criticism of this action at a distance, and it was in this context that he stated the famous phrase "I feign no hypotheses". In modern physics, action at a distance has been completely eliminated, except for subtle effects involving quantum entanglement. However in modern engineering in all practical applications involving the motion of vehicles and satellites, the concept of action at a distance is used extensively. Conservation of energy was discovered nearly two centuries after Newton's lifetime, the long delay occurring because of the difficulty in understanding the role of microscopic and invisible forms of energy such as heat and infra-red light.

Kinetic theory (or the kinetic or kinetic-molecular theory of gases) is the theory that gases are made up of a large number of small particles (atoms or molecules), all of which are in constant, random motion. The rapidly moving particles constantly collide with each other and with the walls of the container. Kinetic theory explains macroscopic properties of gases, such as pressure, temperature, or volume, by considering their molecular composition and motion. Essentially, the theory posits that pressure is due not to static repulsion between molecules, as was Isaac Newton's conjecture, but due to collisions between molecules moving at different velocities. While the particles making up a gas are too small to be visible, the jittering motion of pollen grains or dust particles which can be seen under a microscope, known as Brownian motion, results directly from collisions between the particle and air molecules. This experimental evidence for kinetic theory, pointed out by Albert Einstein in 1905, is generally seen as having confirmed the existence of atoms and molecules.

Postulates
The theory for ideal gases makes the following assumptions:

The gas consists of very small particles, all with non-zero mass. The number of molecules is large such that statistical treatment can be applied.

These molecules are in constant, random motion. The rapidly moving particles constantly collide with the walls of the container. The collisions of gas particles with the walls of the container holding them are perfectly elastic. The interactions among molecules are negligible. They exert no forces on one another except during collisions. The total volume of the individual gas molecules added up is negligible compared to the volume of the container. This is equivalent to stating that the average distance separating the gas particles is large compared to their size. The molecules are perfectly spherical in shape, and elastic in nature. The average kinetic energy of the gas particles depends only on the temperature of the system. Relativistic effects are negligible. Quantum-mechanical effects are negligible. This means that the inter-particle distance is much larger than the thermal de Broglie wavelength and the molecules are treated as classical objects. The time during collision of molecule with the container's wall is negligible as comparable to the time between successive collisions. The equations of motion of the molecules are time-reversible.

More modern developments relax these assumptions and are based on the Boltzmann equation. These can accurately describe the properties of dense gases, because they include the volume of the molecules. The necessary assumptions are the absence of quantum effects, molecular chaos and small gradients in bulk properties. Expansions to higher orders in the density are known as virial expansions. The definitive work is the book by Chapman and Enskog but there have been many modern developments and there is an alternative approach developed by Grad based on moment expansions.[citation needed] In the other limit, for extremely rarefied gases, the gradients in bulk properties are not small compared to the mean free paths. This is known as the Knudsen regime and expansions can be performed in the Knudsen number. The kinetic theory has also been extended to include inelastic collisions in granular matter by Jenkins and others.[citation needed]

[edit] Factors
[edit] Pressure
Pressure is explained by kinetic theory as arising from the force exerted by gas molecules impacting on the walls of the container. Consider a gas of N molecules, each of mass m, enclosed in a cuboidal container of volume V=L3. When a gas molecule collides with the wall of the container perpendicular to the x coordinate axis and bounces off in the opposite direction with the same speed (an elastic collision), then the momentum lost by the particle and gained by the wall is:

where vx is the x-component of the initial velocity of the particle. The particle impacts one specific side wall once every

(where L is the distance between opposite walls). The force due to this particle is:

The total force on the wall is

where the bar denotes an average over the N particles. Because of the Pythagorean , we can rewrite the force as

. This force is exerted on an area L2. Therefore the pressure of the gas is

where V=L3 is the volume of the box. The fraction n=N/V is the number density of the gas (the mass density =nm is less convenient for theoretical derivations on atomic level). Using n, we can rewrite the pressure as

This is a first non-trivial result of the kinetic theory because it relates pressure, a macroscopic property, to the average (translational) kinetic energy per molecule which is a microscopic property.

[edit] Temperature and kinetic energy


From the ideal gas law (1) where is the Boltzmann constant, and the absolute temperature,

and from the above result

we have then the temperature takes the form (2) which leads to the expression of the kinetic energy of a molecule

The kinetic energy of the system is N time that of a molecule The temperature becomes

(3) Eq.(3)1 is one important result of the kinetic theory: The average molecular kinetic energy is proportional to the absolute temperature. From Eq.(1) and Eq.(3)1, we have (4) Thus, the product of pressure and volume per mole is proportional to the average (translational) molecular kinetic energy. Eq.(1) and Eq.(4) are called the "classical results", which could also be derived from statistical mechanics; for more details, see [1]. Since there are degrees of freedom (dofs) in a monoatomic-gas system with particles, the kinetic energy per dof is (5) In the kinetic energy per dof, the constant of proportionality of temperature is 1/2 times Boltzmann constant. This result is related to the equipartition theorem. As noted in the article on heat capacity, diatomic gases should have 7 degrees of freedom, but the lighter gases act as if they have only 5. Thus the kinetic energy per kelvin (monatomic ideal gas) is:

per mole: 12.47 J per molecule: 20.7 yJ = 129 eV

At standard temperature (273.15 K), we get:


per mole: 3406 J per molecule: 5.65 zJ = 35.2 meV

[edit] Number of collisions with wall


One can calculate the number of atomic or molecular collisions with a wall of a container per unit area per unit time. Assuming an ideal gas, a derivation[2] results in an equation for total number of collisions per unit time per area:

[edit] RMS speeds of molecules


From the kinetic energy formula it can be shown that

with v in m/s, T in kelvins, and R is the gas constant. The molar mass is given as kg/mol. The most probable speed is 81.6% of the rms speed, and the mean speeds 92.1% (distribution of speeds).

[edit] History
In 1738 Daniel Bernoulli published Hydrodynamica, which laid the basis for the kinetic theory of gases. In this work, Bernoulli positioned the argument, still used to this day, that gases consist of great numbers of molecules moving in all directions, that their impact on a surface causes the gas pressure that we feel, and that what we experience as heat is simply the kinetic energy of their motion. The theory was not immediately accepted, in part because conservation of energy had not yet been established, and it was not obvious to physicists how the collisions between molecules could be perfectly elastic. Other pioneers of the kinetic theory (which were neglected by their contemporaries) were Mikhail Lomonosov (1747),[3] Georges-Louis Le Sage (ca. 1780, published 1818),[4] John Herapath (1816)[5] and John James Waterston (1843),[6] which connected their research with the development of mechanical explanations of gravitation. In 1856 August Krnig (probably after reading a paper of Waterston) created a simple gas-kinetic model, which only considered the translational motion of the particles. [7] In 1857 Rudolf Clausius, according to his own words independently of Krnig, developed a similar, but much more sophisticated version of the theory which included translational and contrary to Krnig also rotational and vibrational molecular motions. In this same work he introduced the concept of mean free path of a particle. [8] In 1859, after reading a paper by Clausius, James Clerk Maxwell formulated the Maxwell distribution of molecular velocities, which gave the proportion of molecules having a certain velocity in a specific range. This was the first-ever statistical law in physics.[9] In his 1873 thirteen page article 'Molecules', Maxwell states: we are told that an 'atom' is a material point, invested and surrounded by 'potential forces' and that when 'flying molecules' strike against a solid body in constant succession it causes what is called pressure of air and other gases.[10] In 1871, Ludwig Boltzmann generalized Maxwell's achievement and formulated the Maxwell Boltzmann distribution. Also the logarithmic connection between entropy and probability was first stated by him. In the beginning of twentieth century, however, atoms were considered by many physicists to be purely hypothetical constructs, rather than real objects. An important turning point was Albert Einstein's (1905)[11] and Marian Smoluchowski's (1906)[12] papers on Brownian motion, which succeeded in making certain accurate quantitative predictions based on the kinetic theory. Terjmahan indo

Menerjemahkan teks, laman web, dan dokumen


Masukkan teks atau URL laman web, atau unggah dokumen. Terjemahkan dari: Terjemahkan ke: Terjemahan Inggris ke Bahasa Indonesia Teori kinetik (atau kinetik-molekuler atau teori kinetik gas) adalah teori yang menyatakan bahwa gas terdiri dari sejumlah besar partikel kecil (atom atau molekul), semua yang terusmenerus, gerakan acak. Partikel bergerak dengan cepat terus-menerus berbenturan dengan satu sama lain dan dengan dinding wadah. Menjelaskan teori kinetik gas sifat-sifat makroskopik, seperti tekanan, temperatur, atau volume, dengan mempertimbangkan komposisi molekul mereka dan gerak. Pada dasarnya, teori ini berpendapat bahwa tekanan itu disebabkan bukan statis tolakan antara molekul, seperti Isaac Newton's dugaan, namun karena tumbukan antara molekul yang bergerak pada kecepatan yang berbeda.

Sementara partikel yang membentuk gas terlalu kecil untuk dapat dilihat, maka gerakan jittering serbuk sari atau partikel debu yang dapat dilihat di bawah mikroskop, yang dikenal sebagai gerakan Brown, hasil langsung dari tumbukan antara partikel dan molekul udara. Ini bukti eksperimental untuk teori kinetik, ditunjukkan oleh Albert Einstein pada tahun 1905, umumnya dilihat sebagai memiliki mengukuhkan keberadaan atom dan molekul. Postulat Teori gas ideal membuat asumsi sebagai berikut: gas terdiri dari partikel-partikel yang sangat kecil, semua dengan nol non-massa. Jumlah molekul besar sehingga perlakuan statistik dapat diterapkan. Molekul-molekul ini terus-menerus, gerakan acak. Partikel bergerak dengan cepat terusmenerus berbenturan dengan dinding wadah. Para tabrakan partikel gas dengan dinding wadah memegang mereka adalah elastis sempurna. Interaksi antara molekul dapat diabaikan. Mereka tidak mengerahkan pasukan satu sama lain kecuali selama tumbukan. Total volume molekul gas individu dijumlahkan diabaikan dibandingkan dengan volume wadah. Ini sama dengan menyatakan bahwa jarak rata-rata yang memisahkan partikel gas besar dibandingkan dengan ukuran mereka. Molekul-molekul berbentuk bulat sempurna, dan elastis di alam. rata-rata energi kinetik partikel gas hanya bergantung pada suhu sistem. efek relativistik diabaikan. Quantum-efek mekanis dapat diabaikan. Ini berarti bahwa jarak antar partikel jauh lebih besar daripada panjang gelombang de Broglie termal dan molekul diperlakukan sebagai benda klasik. Waktu selama benturan molekul dengan dinding wadah yang diabaikan sebagai sebanding dengan waktu antara tumbukan berturut-turut. The persamaan gerak dari molekul adalah waktu-reversibel. Santai perkembangan lebih modern dan asumsi ini didasarkan pada persamaan Boltzmann. Ini dapat secara akurat menggambarkan sifat-sifat padat gas, karena mereka termasuk volume molekul. Asumsi yang diperlukan adalah tidak adanya efek kuantum, molekul kekacauan dan gradien kecil dalam massal properti. Perluasan perintah yang lebih tinggi dalam kepadatan virial dikenal sebagai ekspansi. Kerja definitif buku oleh Chapman dan Enskog tetapi ada banyak perkembangan modern dan ada pendekatan alternatif yang dikembangkan oleh Grad didasarkan pada saat ekspansi. [Rujukan?] Dalam batasan lain, karena sangat langka gas, maka gradien dalam massal properti tidak kecil jika dibandingkan dengan jalan bebas rata-rata. Hal ini dikenal sebagai rezim dan ekspansi Knudsen dapat dilakukan di nomor Knudsen. Teori kinetik juga telah diperluas untuk mencakup tumbukan inelastis dalam masalah rinci oleh Jenkins dan lain-lain. [Rujukan?] [sunting] Faktor-faktor [sunting] Tekanan Tekanan dijelaskan dengan teori kinetik yang timbul dari gaya yang diberikan oleh molekul gas yang berdampak pada dinding wadah. Pertimbangkan gas N molekul, masing-masing bermassa m, tertutup dalam wadah cuboidal volume V = L3. Ketika molekul gas bertabrakan dengan dinding wadah tegak lurus terhadap sumbu koordinat x dan memantul ke arah berlawanan dengan kecepatan yang sama (tumbukan elastik), maka momentum yang hilang oleh partikel dan dinding yang diperoleh adalah: mana vx adalah x-komponen kecepatan awal partikel. Dampak partikel salah satu dinding sisi tertentu sekali setiap (di mana L adalah jarak antara dinding yang berseberangan). Gaya karena partikel ini adalah: Total gaya pada dinding mana bar menunjukkan rata-rata di atas N partikel. Karena Pythagoras, kita dapat menulis ulang gaya sebagai . Gaya ini diberikan pada suatu daerah L2. Oleh karena itu tekanan gas

di mana V = L3 adalah volume kotak. Fraksi n = N / V adalah jumlah gas kepadatan (densitas massa = nm kurang nyaman bagi teoretis turunan pada tingkat atom). Menggunakan n, kita dapat menulis ulang tekanan sebagai Ini adalah pertama sepele non-akibat dari teori kinetik karena hal ini berkaitan tekanan, makroskopik properti, dengan rata-rata (translasi) energi kinetik per molekul yang merupakan properti mikroskopis. [sunting] Suhu dan energi kinetik Dari hukum gas ideal (1) di mana adalah konstanta Boltzmann, dan suhu absolut, dan dari hasil di atas kita maka suhu mengambil bentuk (2) yang mengarah pada ekspresi dari energi kinetik dari sebuah molekul Energi kinetik dari sistem adalah N waktu yang dari molekul Suhu menjadi (3) Persamaan. (3) 1 adalah salah satu hasil penting dari teori kinetik: Rata-rata energi kinetik molekul sebanding dengan suhu absolut. Dari Persamaan. (1) dan Persamaan. (3) 1, kita telah (4) Dengan demikian, hasil dari tekanan dan volume per mol adalah sebanding dengan ratarata (translasi) energi kinetik molekul. Persamaan. (1) dan Persamaan. (4) disebut sebagai "hasil klasik", yang juga dapat diturunkan dari mekanika statistik; untuk rincian lebih lanjut, lihat [1]. Karena terdapat derajat kebebasan (dofs) dalam sistem monoatomic gas dengan partikel, energi kinetik per dof adalah (5) Dalam energi kinetik per dof, konstanta proporsionalitas suhu adalah 1 / 2 kali Boltzmann konstan. Hasil ini berhubungan dengan teorema equipartition. Seperti tercantum dalam artikel tentang kapasitas panas, gas diatomik harus memiliki 7 derajat kebebasan, tetapi gas yang lebih ringan bertindak seolah-olah mereka hanya 5. Dengan demikian energi kinetik per kelvin (gas ideal monoatomik) adalah: per mol: 12,47 J per molekul: 20,7 j = 129 eV Pada suhu standar (273,15 K), kita mendapatkan: per mol: 3406 J per molekul: 5,65 ZJ = 35,2 MeV [sunting] Jumlah tumbukan dengan dinding Satu dapat menghitung jumlah atom atau molekul tumbukan dengan dinding sebuah wadah per satuan luas per satuan waktu. Dengan asumsi gas ideal, turunan [2] hasil dalam persamaan untuk jumlah tumbukan per satuan waktu per wilayah: [sunting] RMS kecepatan molekul Dari rumus energi kinetik dapat ditunjukkan bahwa dengan v dalam m / s, T di kelvin, dan R adalah konstanta gas. Massa molar diberikan sebagai kg / mol. Kecepatan yang paling mungkin adalah 81,6% dari kecepatan rms dan kecepatan rata-rata 92,1% (distribusi kecepatan).

Benda tegar For analytical convenience, human body segments are considered as rigid bodies. A rigid body is similar to a system of particles in the sense that it is composed of particles. Therefore most of the equations for a system of particles are usable in the dynamics of a rigid body. The main difference is of course that the rigid body is rigid. There is no migration of mass within a rigid body. As a result, the relative positions among the particles composing a rigid body do not change. This will further simplify the equations obtained from a system of particles. The main strategy in analyzing the motion of a rigid body is to split the motion into the linear motion of the CM and the angular motion of the body about its CM. This is because all particles composing the body show the same relative angular motion about the CM. In other words, one can describe the motion of the rigid body as a whole rather than those of the particles individually. The physical characteristics of a rigid body can be described by its inertial properties: mass and moment of inertia. For this reason, a major portion of discussion will be dedicated to the inertial properties of the rigid body: moment of inertia & inertia tensor. The pages included in this section are: Moment of Inertia Calculation of the MOI Inertia Tensor Principal Axes Transformation of the Inertia Tensor Angular Momentum Kinetic Energy Moment of Inertia of a Systems of Particles Newton's first law of motion says "A body maintains the current state of motion unless acted upon by an external force." The measure of the inertia in the linear motion is the mass of the system and its angular counterpart is the so-called moment of inertia. The moment of inertia of a body is not only related to its mass but also the distribution of the mass throughout the body. So two bodies of the same mass may possess different moments of inertia. A rigid body can be considered as a system of particles in which the relative positions of the particles do not change. The moment of inertia of a single particle (I) can be expressed as [1] where m = the mass of the particle, and r = the shortest distance from the axis of rotation to the particle (Figure 1).

Figure 1 As shown in [1], moment of inertia is equal to mass times square of the distance and it is also referred to as the second mass moment. Mass times distance, mr, is called as the first mass moment. This concept of first mass moment is normally used in deriving the center of mass of a system of particles or a rigid body. See Center of Mass-System of Particles for the details. Expanding [1] for a system of particles:

[2] Top

Moment of Inertia of a Rigid Body Based on [2], one can obtain the moment of inertia of a rigid by shown in Figure 2:

Figure 2

[3] where ri = the position of particle i, and n = the unit vector of the axis of rotation. Note here that the axis of rotation passes through the local reference frame, the OXYZ system. Let

[4] and

[5] where cos, cos & cos = the three direction cosines of vector n to the XYZ system. Substituting [4] & [5] into [3] leads to

[6] where

[7] Ixx, Iyy & Izz are called the moments of inertia while Ixy, Iyx, Iyz, Izy, Izx, & Ixz are the products of inertia. For a rigid body, the relative position of the particles do not change and one can write [7] as:

[8] When the shape and the density distribution of the rigid body is precisely known, one can use [8] to compute the moments and products of inertia. (See BSP Equations for the MOI equations of the typical geometric shapes commonly used in human body modeling.) Otherwise, it is difficult to compute them through integration. Rather, the moment of inertia must be measured directly from the object. See Measuring MOI for the details. Top

Ellipsoid of Inertia The moments and products of inertia shown in [7] and [8] are basically specific to the local reference frame defined and reflect the mass distribution within the body in relation to the local reference frame. As shown in [6], the actual moment of inertia of a rigid body about an axis of rotation is a function of not only the moments and products of inertia for a given reference frame but also the orientation of the axis of rotation, , & . Thus, it would be more accurate to say

that the moment of inertia of a rigid body reflects the mass distribution within the body with respect to the axis of rotation. As the axis of rotation changes, so does the moment of inertia. To show this point clearly, let

[9] Substituting [9] into [6] yields [10] Interestingly, [10] suffices the general form of the ellipsoid with its center at the origin of the reference frame. When Ixy = Iyz = Izx = 0, the ellipsoid defined by [10] definitely becomes symmetric about the three axes. Since

[11] the distance from the center of the ellipsoid to the surface is 1 divided by the square root of the moment of inertia of the rigid body for a given orientation, , & . The ellipsoid defined by [10] is called the ellipsoid of inertia since it describes the moment of inertia of an object as a function of the orientation of the axis of rotation. Top Calculation of the MOI In computation of the moment of inertia, one can replace the summation shown in [2] of Inertia Tensor by an integration over the body: [1] where r = the perpendicular distance from the particle to the axis of rotation, and dm = the mass of the particle which is a function of the density. Thin Rod Let's now apply [1] to a thin uniform rod shown in Figure 1. The MOI of the rod about the Y axis is

Figure 1

[2] since

[3] where = the density of the rod, m = the mass of the rod, and L = the length of the rod.

Circular Ring The moment of inertia of the uniform circular ring shown in Figure 2 about the Z axis (the symmetry axis) is

Figure 2

[4] since

[5] where dl = the length of the arc formed by d . [5] is also applicable to a circular cylinder.

Circular Disc A uniform circular disc of radius R can be considered as a cascade of uniform circular rings as shown in Figure 3. Thus, from [5], the moment of inertia about the Z axis (the symmetry axis) becomes

Figure 3

[6] since [7] [6] can be used in computing the MOI of a circular bar about its longitudinal axis as well.

Sphere A uniform sphere can be considered as a cascade of uniform circular discs as shown in Figure 4. From [6], the moment of inertia about the vertical axis (Z axis) is

Figure 4

[8] since

[9] See BSP Equations for the moment-of-inertia equations for the geometric shapes commonly used in the human body modeling. Top

Perpendicular-Axis Theorem Now, let's go back to the uniform circular disc (Figure 3). The moment of inertia of a uniform circular disc about its perpendicular axis (Z axis) can be expressed as

[10] since

[11] In other words, the MOI about the Z axis is equal to sum of those about the X and Y axes. [10] is true for any rigid lamina: the MOI of any rigid lamina about an axis normal to the lamina plane is equal to the sum of the MOIs about any two perpendicular axes lying on the plane and passing through the normal axis. This is the so-called perpendicular-axis theorem. Since the circular disc has symmetric shape, [12] and, from [6]:

[13] One can directly obtain [13] from [11] or

[14] [2] can be used in further developing [14] to obtain [13]. See BSP Equations for this approach. Top

Parallel-Axis Theorem Now, let's compute the MOI of a uniform circular column. Circular bar can be regarded as a cascade of circular discs as shown in Figure 5. From [13]:

Figure 5

[15] since

[16] where mdisc = mass of the circular disc, and Iy'y'(disc) = the MOI of the disc about the Y' axis. [16] basically says that the MOI of a circular plate about the Y axis is equal to the sum of the MOI about the parallel axis on the disc (Y') and the mass of the disc times square of the distance between the two axes. This is the socalled parallel-axis theorem. The MOI of the circular column is therefore

[17] since [18] Comparing [18] with [2], one can clearly see the difference in the MOI between a thin rod and a thick rod (circular column). Similarly, [13] and [17] shows the difference in the MOI between a circular disc and a thick circular plate (circular column). Top

Physical Pendulum & Direct Measurement Unfortunately, the integration approach is possible only when the body has a known geometric shape. In the mathematical human body models such as Hanavan (1964) and Yeadon (1990), it is assumed that the body segments show a group of geometric shapes such as ellipsoid of revolution, elliptical solids, and stadium solids. See BSP Equations for the details. If the body has a irregular shape, the integration approach has not much use and a direct measurement must be attempted. Figure 6 shows a body with irregular shape which is rotating freely about an axis passing through its one end. The X axis is the axis of rotation, thus, the center of mass (CM) of the body moves within the YZ plane.

Figure 6 The torque produced by the weight of the body about the X axis is then [19] where Tx = the torque about the X axis, Ixx = MOI of the body about the X axis, = angular acceleration, m = the mass of the body, g = the gravitational acceleration (9.81 m/s2), and L = the distance between the axis of rotation to the body's CM. For a small , [20] and, from [19], [21] Solving [21] for , one obtains

[22] where o = the amplitude, f = the frequency of the pendulum, = the phase angle, T = the period of the pendulum. As shown in [22], the MOI of the body about the X axis, after all, can be computed from the period of a small pendulum motion of the body. The MOI about the parallel axis, which passes through the CM of the body, can be also computed based on the parallel-axis theorem: [23] See Chandler et al. (1975) for an example of this approach. Angular Momentum of a Rigid Body Angular momentum of a rigid body (Figure 1) can be obtained from the sum of the angular momentums of the particles forming the body:

[1]

Figure 1

where ri = the position vector of particle i, and = the angular velocity vector of the rigid body. Now, let

[2]

[3] Note here that a local reference frame, the XYZ system, is defined and fixed to the body at O and ri & are described in this frame. Substituting [2] & [3] into [1]:

[4] Top

Inertia Tensor Let

[5] Substituting [5] into [4]:

[6]

[7] where I = the inertia tensor. The angular momentum of a rigid body rotating about an axis passing through the origin of the local reference frame is in fact the product of the inertia tensor of the object and the angular velocity. The diagonal elements in the inertia tensor shown in [7], Ixx, Iyy & Izz, are called the moments of inertia while the rest of the elements are called the products of inertia. Also see Moment of Inertia & Ellipsoid of Inertia for more details of the moments and products of inertia. As shown in [7], the inertia tensor is symmetric. The 3 x 3 matrix in [7] suffices the requirements of a tensor of the 2nd rank:

[8] where i, j, k & l = 1 to 3, tij = an element of the orthogonal transformation matrix, and I'ij = an element of the transformed inertia tensor. This property is explained in detail in Transformation of the Inertia Tensor. In fact, a scalar is a tensor of zero rank: [9] where S = a scalar, S' = the transformed scalar. In addition, a vector is a tensor of the first rank since its transformation follows

[10] [10] is identical to [2] shown in Transformation Matrix. A 2-dimensional symmetric matrix is not necessarily a tensor of the 2nd rank. It must suffice [8] to be a tensor. Principal Moments of inertia As shown in [6] in Inertia Tensor, the angular momentum of a rigid body with respect to the origin of the local reference frame is expressed as

[1] If, by chance, all the off-diagonal terms of the inertia tensor shown in [1] become zero, [1] can be further simplified to

[2] This can happen when one aligns the axes of the local reference frame in such a way that the mass of the body evenly distribute around the axes, thus, the product-of-inertia terms all vanish. The non-zero diagonal terms of the inertia tensor shown in [2] are called the principal moments of inertia of the object. Top

Principal Axes As shown in [1], there is no guarantee that the angular momentum vector has the same direction to that of the angular velocity vector. This causes a problem: if the direction of the angular momentum keeps changing, it develops a torque which eventually forces the axis of rotation to move. This is the main reason that causes wearing and vibration in machinery with rotating parts. But in some special cases, the following condition may hold so that the angular momentum and velocity vectors show the same direction:

[3] where I = the equivalent scalar moment of inertia of the body about the axis of rotation. Any axis of rotation of the body which suffices [3] is called a principal axis. There are a group of principal axes (theoretically 3) in a three-dimensional body. For example, there are three perpendicular principal axes for the system shown in Figure 1.

Figure 1 [3] basically says that the inertia tensor can be replaced with a single scalar moment of inertia when the axis of rotation is a principal axis. Top

Diagonalization of the Inertia Tensor From [3]:

[4] Or [4] can be simplified to [5] where 1 = the identity matrix. I shown in [5] is called an eigenvalue while is the eigenvector. [5] is the eigenvalue equation. In order for [5] to have a nontrivial solution the determinant of the coefficients should vanish:

[6] [6] leads to the secular equation which is basically cubic, thus provides three roots (eigenvalues): I1, I2 & I3. Each root corresponds to a moment of inertia about a principal axis. In fact the three roots are the principal moments of inertia of the rigid body introduced in [2]:

[7] Once the eigenvalues are known, the principal axes can be computed. Let

[8] where n = the unit vector of the principal axis, thus, [9] From [4] & [8]:

[10] For each eigenvalue, one can compute the corresponding nx, ny & nz from [9] & [10]. One must pay attention to the direction of the eigenvector in this process.

In motion analysis, the principal moments of inertia can be obtained from the inertial properties of the body segments. I1, I2 & I3 of each segment are generally known. The data are available in the form of the radius-of-gyration ratios (ratio of the radius of gyration to the segment length), regression equations, and scaling coefficients. One can also compute the principal moments of inertia of the body segments through modeling using some geometric shapes. See Individualized BSP Estimation for the details. Top Angular Momentum of a Rigid Body As shown in [6] in Inertis Tensor, the angular momentum of a rigid body rotating about an axis passing through the origin of the local reference frame (frame A) is [1] Now, let's transforms this angular momentum vector to another reference frame (frame B): [2] Combination of [1] & [2] leads to

[3] Top

Rotation [3] shows that the angular momentum of the object described in frame B is equal to I' times the angular velocity described in frame B. The axis transformation does not alter the nature of any mechanical quantity, but describes it in a different perspective. Since both the angular momentum & the angular velocity vectors are already described in frame B, I' must be the inertia tensor of the object described in frame B: [4] This type of transformation is called the similarity transformation. Now, let A be a local reference frame (frame L) and B be the global (inertial) frame. Then, the inertia tensor of a body segment described in the global coordinates is

[5] In motion analysis, one can compute the transformation matrix from the global frame to a segmental reference frame based on the marker data, while the inertia

tensor is typically first described in the corresponding segmental reference frame. [5] can be used to compute the inertia tensor described in the global frame. Top

Tensor of the 2nd Rank From [4]:

[6] where i, j, k, & m = 1 to 3. Or simplifying [6], one can obtain

[7] [7] is identical to [8] in Inertia Tensor. [7] is the requirement of a tensor of the 2nd rank. A scalar is a tensor or zero rank, while a vector is a tensor of the 1st rank. Top

Translation Figure 1 shows two reference frames: the OXYZ system & the O'X'Y'Z' system. The XYZ system is the local reference frame fixed to the body with its origin at the body's center of mass (CM). The X'Y'Z' system, on the other hand, is parallel to the XYZ system with different origin. From [4] of Inertia Tensor, the inertia tensor of the body about the X'Y'Z' system is

Figure 1

[8] Now, let [9] and therefore,

[10] Substituting [10] into [8] gives

[11] since

[12] where m = the mass of the body. [12] basically means the sum of the first mass moments about the body's CM is always 0. With [11], [8] is reduced to

[13] where

[14] ICM is the inertia tensor of the body about its CM while It is the additional MOI due to the translation of the reference frame. Matrix 1 shown in [14] is the identity matrix whose diagonal elements are all 1 while the off-diagonal elements are 0. As shown in [13], the moment of inertia of an object about a reference frame parallel to the local reference frame fixed to the body is equal to the sum of the inertia tensor of the body and the additional MOI due to translation of the origin of the frame. Here is an example. Let

[15] where Ixx, Iyy & Izz = the principal moments of inertia of the body. In other words, the X, Y & Z axes shown in Figure 1 are the principal axes. See Principal Axes for the details of the principal moments of inertia and axes. If the Z' axis shown in Figure 1 is the axis of rotation or:

[16] the angular momentum of the body about the axis of rotation becomes

[17] If we further assume that zo = 0, in other words, both the XYZ and the X'Y'Z' systems share the XY plane,

[18] where I' = the moment of inertia about the axis of rotation (Z' axis), and d = the distance between the axis of rotation & the parallel axis passing through the body's CM (Z axis). Izz in [18] is in fact the moment of inertia of the object about the Z axis. [18] may be generalized to the following form: [19] where I' = the moment of inertia of the body about the axis of rotation, ICG = the moment of inertia of the body's principal axis which is parallel to the axis of rotation, m = mass of the body, and d = the distance between the two axes. [19] is the so-called the parallel-axis theorem. Top Angular Momentum of a Rigid Body

Figure 1 A rigid body is both rotating about its center of mass and translating with respect to the origin of the reference frame, O shown in Figure 1. From [1] of Angular Momentum - System of Particles and Figure 1, the angular momentum of a particle composing the rigid body is

[1] where m = the mass of the particle, rCM = the position of the body's CM, vCM = the velocity of the body's CM, r = the position of the particle, r' = the relative position of the particle to the body's CM, v = the velocity of the particle, and v' = the relative velocity of the particle to the body's CM. Since all the particles in the body experience the same angular velocity, [1] can be rewritten to [2] since, from Figure 1, [3] where = the angular velocity of the body. The angular momentum of the body is the sum of the angular momentums of the particles composing the body:

[4] where M = the mass of the rigid body. Top

Remote & Local Terms From [4] of Inertia Tensor and [4]:

[5] where pCM = the momentum of the body, ICM = the inertia tensor of the body about its CM, HCM = the angular momentum of the body due to the motion of the CM, and H' = the angular momentum of the body due to its rotation about the CM. As shown in [5], the angular momentum of a rigid body can be reduced to two distinct terms: the angular momentum due to the translation of the body's CM (HCM), and the angular momentum due to the rotation of the body about its CM (H'). The first term in [5] is called the remote term, while the second is the local term of the body's angular momentum:

[6] Top

Kinetic Energy ]

Vous aimerez peut-être aussi