Search This Blog

Tuesday, February 10, 2015

The Planetary Society


From Wikipedia, the free encyclopedia

The Planetary Society
Type NGO and Non-profit foundation
Location
Fields Space advocacy
Members
40,000+
Key people
Carl Sagan, Bruce C. Murray, Louis Friedman, Bill Nye, Neil deGrasse Tyson
Mission To inspire the people of Earth to explore other worlds, understand our own, and seek life elsewhere.
Website planetary.org

Planetary Society founders. 1980 photo.

The Planetary Society is an American-based non-government, nonprofit organization; anyone can join. It is involved in research and engineering projects related to astronomy, planetary science, exploration, public outreach, and political advocacy. It was founded in 1980 by Carl Sagan, Bruce Murray, and Louis Friedman,[1] and has over 40,000 members[citation needed] from more than one hundred countries around the world.

The Society is dedicated to the exploration of the Solar System, the search for Near Earth Objects, and the search for extraterrestrial life.[2] The society's mission is stated as: "To empower the world's citizens to advance space science and exploration".[3]

In addition to public outreach, the Planetary Society also sponsors novel and innovative projects that will "seed" further exploration. In June 2005, the Society launched the Cosmos 1 craft to test the feasibility of solar sailing, but the rocket failed shortly after liftoff.[4][5]

The Planetary Society runs many programs. Two of the highest profile programs are Lightsail and LIFE (Living Interplanetary Flight Experiment.) Lightsail is a series of three solar sail experiments.[6] LightSail-1 is expected to piggyback on a future NASA mission.[7]

LIFE was a two-part program designed to test the ability of microorganisms to survive in space.[8] The first phase flew on shuttle Endeavor's final flight in 2011.[9] The second phase rode on Russia's Fobos-Grunt mission, which attempted to go to Mars' moon Phobos and back but failed to escape earth orbit. [10]

In addition to its projects, The Planetary Society is also a strong advocate for space funding and missions of exploration within NASA. They actively lobby Congress and engage their membership in the United States to write and call their representatives in support of NASA funding.

History

The Planetary Society was founded in 1980 by Carl Sagan, Bruce Murray, and Louis Friedman as a champion of public support of space exploration and the search for extraterrestrial life. Until his death in 1996, the Society was actively led by Sagan, who used his celebrity and political clout to influence the political climate of the time, including protecting SETI in 1981 from congressional cancellation. Throughout the 1980s and 1990s, the Society pushed its scientific and technologic agenda, which led to an increased interest in rover-based planetary exploration and NASA's New Horizons mission to Pluto.

In addition to its political affairs, the Society has created a number of space related projects and programs. The SETI program began with Paul Horowitz's Suitcase SETI and has grown to encompass searches in radio and optical wavelengths from the northern and southern hemispheres of the Earth. SETI@home, the largest distributed computing experiment on Earth, is perhaps the Society's best-known SETI project. Other projects include the development of the Mars Microphone instrument which flew on the failed Mars Polar Lander project, as well as the LightSail-1 project, a solar sail project to determine if space travel is possible by using only sunlight.

Program summary

The Planetary Society currently runs seven different program areas with a number of programs in each area:

Organization

The Planetary Society is governed by a 17-member volunteer board of directors chosen for their passion about and knowledge of space exploration. The Board has a chairman, President, and Vice President and an Executive Committee, and normally meets twice per year to set the Society's policies and future directions. Nominations are sought and considered periodically from a variety of sources, including from members of the Board and Advisory Council, Society Members, staff, and experts in the space community.[11] On June 7, 2010, the Society announced that famed American science educator Bill Nye would become the new executive director of the society.[12]

Members

The Planetary Society's current board of directors includes the following:[11]
Notable members of its Advisory Council include:[13]

Science and Technology

The Planetary Society sponsors science and technology projects to seed further exploration. All of these projects are funded by the Society's members and donors. Some projects include:
A member's donation of $4.2 million in 2014 will be used by the Society to further their research into solar sails and asteroid tracking.[15]

The Planetary Report

The Planetary Report is the quarterly internationally recognized flagship magazine of The Planetary Society, featuring articles and full-color photos to provide comprehensive coverage of discoveries on Earth and other planets. It went from bimonthly to quarterly with the June (summer solstice) 2011 issue.

This magazine reaches over 40,000 members of The Planetary Society all over the world, with news about planetary missions, spacefaring nations, space explorers, planetary science controversies, and the latest findings in humankind's exploration of the solar system.

Planetary Radio

The Planetary Society also produces Planetary Radio, a weekly 30-minute radio program and podcast hosted and produced by Mat Kaplan. The show's programming consists mostly of interviews and telephone-based conversations with scientists, engineers, project managers, artists, writers, astronauts, and many other professionals who can provide some insight or perspective into the current state of space exploration.

Only A Small Fraction Of Cancers Are Unpreventable

Posted on  

shutterstock_breast cancer women
CREDIT: Shutterstock

A significant number of Americans believe they have no power in preventing cancer, even though factors that remain out of their control — like genetics and family historyaccount for less than 20 percent of all cancers.

A recent Cancer Risk Awareness Survey confirms this pervasive information gap. According to the survey, released by the American Institute for Cancer Research earlier this month, 17 percent of participants believed that there was nothing they could to reduce their risk of getting cancer. Researchers also found that nearly 80 percent of respondents remained unaware that genetics’ impact on their cancer risk paled in comparison to poor diet and lack of exercise.

Respondents also expressed concern about food additives, genetically modified foods, stress, and hormones in beef as key causes of cancer — even though research does not support the idea that those are key factors in the development of cancer.

“Instead of focusing on factors that you can’t always control, we want Americans to learn more about the factors that you can and do control, every day,” Alice Bender, the organization’s associate director of nutrition programs, said in a press statement.

Cancer — which counts as the second leading cause of death worldwide — often involves abnormal cell growth that spreads throughout the body, eventually crippling and killing the afflicted. While death rates have been on the decline in recent decades, it still poses a significant threat among smokers, the obese, and those with poor diets. Making matters worse, the number of Americans who know about the benefits of diets high in fruits and vegetables have declined by 10 percentage points since 2009. And more than half of Americans remain unaware about the link between alcohol and cancer.

Recent information about the link between cancer and “bad luck” can further complicate matters to convince at-risk Americans to make lifestyle changes. A study published in the Science Journal earlier this year confirmed that 22 of 31 cancers examined occurred as a result of random gene mutations that occur when stem cell divide. While the groundbreaking research could help members of the medical profession improve their screenings for potential mutations, it could also fuel the misconception that people don’t need to take their lives into their own hands.

“The vast majority of research does not show that cancer is the result of bad luck,” Dr. David J. Hunter, a professor of cancer prevention and dean for academic affairs at the Harvard School of Public Health, wrote in an op-ed in the Boston Globe in response to the news coverage of that study.
“The real question is whether we are going to take cancer prevention seriously by ensuring that populations in the United States and around the world have accurate information, are vaccinated against or treated for cancer-causing infections, and are less likely to engage in cancer-risk behaviors.”

The Centers for Disease Control and Prevention recommends that concerned family members, friends, and neighbors of people at risk of developing cancer should use community resources — including low-cost cancer screenings, workshops, and support groups — to develop strategies that better inform people about the risks associate with unhealthy habits and sedentary living.

Other factors that the government agency suggests that outreach coordinators keep in mind when leading awareness campaigns in communities include income levels, language and cultural barriers, and cultural beliefs about cancer prevention. Those obstacles have often proven detrimental to African American and Latino populations.

That knowledge, or at least an acknowledgment, that lifestyle choices matter in reducing one’s cancer risk can do wonders for many Americans. Nearly a third of the most common cancers in the United States could be prevented if American exercise more, weighed less, and made healthier lifestyle choices, according to a another study conducted by the American Cancer Research Institute. The chances of developing malignant tumors also fell by 50 percent with avoidance of harmful ultraviolet rays. Heeding this advice could potentially cut annual medical costs by $20 billion.

Tensor


From Wikipedia, the free encyclopedia

Cauchy stress tensor, a second-order tensor. The tensor's components, in a three-dimensional Cartesian coordinate system, form the matrix

\begin{align}
\sigma & = \begin{bmatrix}\mathbf{T}^{(\mathbf{e}_1)} \mathbf{T}^{(\mathbf{e}_2)} \mathbf{T}^{(\mathbf{e}_3)} \\ \end{bmatrix} \\
& = \begin{bmatrix} \sigma_{11} & \sigma_{12} & \sigma_{13} \\ \sigma_{21} & \sigma_{22} & \sigma_{23} \\ \sigma_{31} & \sigma_{32} & \sigma_{33} \end{bmatrix}\\
\end{align}

whose columns are the stresses (forces per unit area) acting on the e1, e2, and e3 faces of the cube.

Tensors are geometric objects that describe linear relations between vectors, scalars, and other tensors. Elementary examples of such relations include the dot product, the cross product, and linear maps. Vectors and scalars themselves are also tensors. A tensor can be represented as a multi-dimensional array of numerical values. The order (also degree) of a tensor is the dimensionality of the array needed to represent it, or equivalently, the number of indices needed to label a component of that array. For example, a linear map can be represented by a matrix (a 2-dimensional array) and therefore is a 2nd-order tensor. A vector can be represented as a 1-dimensional array and is a 1st-order tensor. Scalars are single numbers and are thus 0th-order tensors. The dimensionality of the array should not be confused with the dimension of the underlying vector space.

Tensors are used to represent correspondences between sets of geometric vectors; for applications in engineering and Newtonian physics these are normally Euclidean vectors. For example, the Cauchy stress tensor T takes a direction v as input and produces the stress T(v) on the surface normal to this vector for output thus expressing a relationship between these two vectors, shown in the figure (right).

Because they express a relationship between vectors, tensors themselves must be independent of a particular choice of coordinate system. Finding the representation of a tensor in terms of a coordinate basis results in an organized multidimensional array representing the tensor in that basis or frame of reference. The coordinate independence of a tensor then takes the form of a "covariant" transformation law that relates the array computed in one coordinate system to that computed in another one. The precise form of the transformation law determines the type (or valence) of the tensor. The tensor type is a pair of natural numbers (n, m) where n is the number of contravariant indices and m is the number of covariant indices. The total order of a tensor is the sum of these two numbers.

Tensors are important in physics because they provide a concise mathematical framework for formulating and solving physics problems in areas such as elasticity, fluid mechanics, and general relativity. Tensors were first conceived by Tullio Levi-Civita and Gregorio Ricci-Curbastro, who continued the earlier work of Bernhard Riemann and Elwin Bruno Christoffel and others, as part of the absolute differential calculus. The concept enabled an alternative formulation of the intrinsic differential geometry of a manifold in the form of the Riemann curvature tensor.[1]

Definition

There are several approaches to defining tensors. Although seemingly different, the approaches just describe the same geometric concept using different languages and at different levels of abstraction.

As multidimensional arrays

Just as a vector with respect to a given basis is represented by an array of one dimension, any tensor with respect to a basis is represented by a multidimensional array. The numbers in the array are known as the scalar components of the tensor or simply its components. They are denoted by indices giving their position in the array, as subscripts and superscripts, after the symbolic name of the tensor. In most cases, the indices of a tensor are either covariant or contravariant, designated by subscript or superscript, respectively. The total number of indices required to uniquely select each component is equal to the dimension of the array, and is called the order, degree or rank of the tensor.[Note 1] For example, the entries of an order 2 tensor T would be denoted Tij, Ti j, Tij, or Tij, where i and j are indices running from 1 to the dimension of the related vector space.[Note 2] When the basis and its dual coincide (i.e. for an orthonormal basis), the distinction between contravariant and covariant indices may be ignored; in these cases Tij or Tij could be used interchangeably.[Note 3]

Just as the components of a vector change when we change the basis of the vector space, the entries of a tensor also change under such a transformation. Each tensor comes equipped with a transformation law that details how the components of the tensor respond to a change of basis. The components of a vector can respond in two distinct ways to a change of basis (see covariance and contravariance of vectors), where the new basis vectors \mathbf{\hat{e}}_i are expressed in terms of the old basis vectors \mathbf{e}_j as,
\mathbf{\hat{e}}_i = \sum_j R^j_i \mathbf{e}_j = R^j_i \mathbf{e}_j,
where Ri j is a matrix and in the second expression the summation sign was suppressed (a notational convenience introduced by Einstein that will be used throughout this article).[Note 4] The components, vi, of a regular (or column) vector, v, transform with the inverse of the matrix R,
\hat{v}^i = (R^{-1})^i_j v^j,
where the hat denotes the components in the new basis. While the components, wi, of a covector (or row vector), w transform with the matrix R itself,
\hat{w}_i = R_i^j w_j.
The components of a tensor transform in a similar manner with a transformation matrix for each index. If an index transforms like a vector with the inverse of the basis transformation, it is called contravariant and is traditionally denoted with an upper index, while an index that transforms with the basis transformation itself is called covariant and is denoted with a lower index. The transformation law for an order-m tensor with n contravariant indices and mn covariant indices is thus given as,
\hat{T}^{i_1,\ldots,i_n}_{i_{n+1},\ldots,i_m}= (R^{-1})^{i_1}_{j_1}\cdots(R^{-1})^{i_n}_{j_n} R^{j_{n+1}}_{i_{n+1}}\cdots R^{j_{m}}_{i_{m}}T^{j_1,\ldots,j_n}_{j_{n+1},\ldots,j_m}.
Such a tensor is said to be of order or type (n, mn).[Note 5] This discussion motivates the following formal definition:[2]
Definition. A tensor of type (n, mn) is an assignment of a multidimensional array
T^{i_1\dots i_n}_{i_{n+1}\dots i_m}[\mathbf{f}]
to each basis f = (e1,...,eN) such that, if we apply the change of basis
\mathbf{f}\mapsto \mathbf{f}\cdot R = \left( R_1^i \mathbf{e}_i, \dots, R_N^i\mathbf{e}_i\right)
then the multidimensional array obeys the transformation law
T^{i_1\dots i_n}_{i_{n+1}\dots i_m}[\mathbf{f}\cdot R] = (R^{-1})^{i_1}_{j_1}\cdots(R^{-1})^{i_n}_{j_n} R^{j_{n+1}}_{i_{n+1}}\cdots R^{j_{m}}_{i_{m}}T^{j_1,\ldots,j_n}_{j_{n+1},\ldots,j_m}[\mathbf{f}].

The definition of a tensor as a multidimensional array satisfying a transformation law traces back to the work of Ricci.[1] Nowadays, this definition is still used in some physics and engineering text books.[3][4]

Tensor fields

In many applications, especially in differential geometry and physics, it is natural to consider a tensor with components that are functions of the point in a space. This was the setting of Ricci's original work. In modern mathematical terminology such an object is called a tensor field, often referred to simply as a tensor.[1]
In this context, a coordinate basis is often chosen for the tangent vector space. The transformation law may then be expressed in terms of partial derivatives of the coordinate functions,
\bar{x}_i(x_1,\ldots,x_k),
defining a coordinate transformation,[1]
\hat{T}^{i_1\dots i_n}_{i_{n+1}\dots i_m}(\bar{x}_1,\ldots,\bar{x}_k) =
\frac{\partial \bar{x}^{i_1}}{\partial x^{j_1}}
\cdots
\frac{\partial \bar{x}^{i_n}}{\partial x^{j_n}}
\frac{\partial x^{j_{n+1}}}{\partial \bar{x}^{i_{n+1}}}
\cdots
\frac{\partial x^{j_m}}{\partial \bar{x}^{i_m}}
T^{j_1\dots j_n}_{j_{n+1}\dots j_m}(x_1,\ldots,x_k).

As multilinear maps

A downside to the definition of a tensor using the multidimensional array approach is that it is not apparent from the definition that the defined object is indeed basis independent, as is expected from an intrinsically geometric object. Although it is possible to show that transformation laws indeed ensure independence from the basis, sometimes a more intrinsic definition is preferred. One approach is to define a tensor as a multilinear map. In that approach a type (n, m) tensor T is defined as a map,
 T: \underbrace{ V^* \times\dots\times V^*}_{n \text{ copies}} \times \underbrace{ V \times\dots\times V}_{m \text{ copies}} \rightarrow \mathbf{R},
where V is a (finite-dimensional) vector space and V* is the corresponding dual space of covectors, which is linear in each of its arguments.

By applying a multilinear map T of type (n, m) to a basis {ej} for V and a canonical cobasis {εi} for V*,
T^{i_1\dots i_n}_{j_1\dots j_m} \equiv T(\mathbf{\varepsilon}^{i_1},\ldots,\mathbf{\varepsilon}^{i_n},\mathbf{e}_{j_1},\ldots,\mathbf{e}_{j_m}),
an (n+m)-dimensional array of components can be obtained. A different choice of basis will yield different components. But, because T is linear in all of its arguments, the components satisfy the tensor transformation law used in the multilinear array definition. The multidimensional array of components of T thus form a tensor according to that definition. Moreover, such an array can be realised as the components of some multilinear map T. This motivates viewing multilinear maps as the intrinsic objects underlying tensors.

Using tensor products

For some mathematical applications, a more abstract approach is sometimes useful. This can be achieved by defining tensors in terms of elements of tensor products of vector spaces, which in turn are defined through a universal property. A type (n, m) tensor is defined in this context as an element of the tensor product of vector spaces,[5]
 T\in \underbrace{V \otimes\dots\otimes V}_{n \text{ copies}} \otimes \underbrace{V^* \otimes\dots\otimes V^*}_{m \text{ copies}}.
If vi is a basis of V and wj is a basis of W, then the tensor product VW has a natural basis viwj.
The components of a tensor T are the coefficients of the tensor with respect to the basis obtained from a basis {ei} for V and its dual {εj}, i.e.
T = T^{i_1\dots i_n}_{j_1\dots j_m}\; \mathbf{e}_{i_1}\otimes\cdots\otimes \mathbf{e}_{i_n}\otimes \mathbf{\varepsilon}^{j_1}\otimes\cdots\otimes \mathbf{\varepsilon}^{j_m}.
Using the properties of the tensor product, it can be shown that these components satisfy the transformation law for a type (m, n) tensor. Moreover, the universal property of the tensor product gives a 1-to-1 correspondence between tensors defined in this way and tensors defined as multilinear maps.

Tensors in infinite dimensions

This discussion of tensors so far assumes finite dimensionality of the spaces involved. All of this can be generalized, essentially without modification, to vector bundles or coherent sheaves.[6] For infinite-dimensional vector spaces, inequivalent topologies lead to inequivalent notions of tensor, and these various isomorphisms may or may not hold depending on what exactly is meant by a tensor (see topological tensor product). In some applications, it is the tensor product of Hilbert spaces that is intended, whose properties are the most similar to the finite-dimensional case. A more modern view is that it is the tensors' structure as a symmetric monoidal category that encodes their most important properties, rather than the specific models of those categories.

Examples

This table shows important examples of tensors, including both tensors on vector spaces and tensor fields on manifolds. The tensors are classified according to their type (n, m), where n is the number of contravariant indices, m is the number of covariant indices, and n + m gives the total order of the tensor. For example, a bilinear form is the same thing as a (0, 2)-tensor; an inner product is an example of a (0, 2)-tensor, but not all (0, 2)-tensors are inner products. In the (0, M)-entry of the table, M denotes the dimensionality of the underlying vector space or manifold because for each dimension of the space, a separate index is needed to select that dimension to get a maximally covariant antisymmetric tensor.
n, m n = 0 n = 1 n = 2 ... n ...
m = 0 scalar, e.g. scalar curvature vector (e.g. direction vector) inverse metric tensor, bivector (e.g., a Poisson structure) n-vector, a sum of n-blades
m = 1 covector, linear functional, 1-form (e.g. gradient of a scalar field) linear transformation,[7] Kronecker delta
m = 2 bilinear form, e.g. inner product, metric tensor, Ricci curvature, 2-form, symplectic form e.g. cross product in three dimensions e.g. elasticity tensor
m = 3 e.g. 3-form e.g. Riemann curvature tensor
...
m = M e.g. M-form i.e. volume form
...
Raising an index on an (n, m)-tensor produces an (n + 1, m − 1)-tensor; this can be visualized as moving diagonally up and to the right on the table. Symmetrically, lowering an index can be visualized as moving diagonally down and to the left on the table. Contraction of an upper with a lower index of an (n, m)-tensor produces an (n − 1, m − 1)-tensor; this can be visualized as moving diagonally up and to the left on the table.

Orientation defined by an ordered set of vectors.
Reversed orientation corresponds to negating the exterior product.

Geometric interpretation of grade n elements in a real exterior algebra for n = 0 (signed point), 1 (directed line segment, or vector), 2 (oriented plane element), 3 (oriented volume). The exterior product of n vectors can be visualized as any n-dimensional shape (e.g. n-parallelotope, n-ellipsoid); with magnitude (hypervolume), and orientation defined by that on its n − 1-dimensional boundary and on which side the interior is.[8][9]

Notation

Ricci calculus

Ricci calculus is the modern formalism and notation for tensor indices: indicating inner and outer products, covariance and contravariance, summations of tensor components, symmetry and antisymmetry, and partial and covariant derivatives.

Einstein summation convention

The Einstein summation convention dispenses with writing summation signs, leaving the summation implicit. Any repeated index symbol is summed over: if the index i is used twice in a given term of a tensor expression, it means that the term is to be summed for all i. Several distinct pairs of indices may be summed this way.

Penrose graphical notation

Penrose graphical notation is a diagrammatic notation which replaces the symbols for tensors with shapes, and their indices by lines and curves. It is independent of basis elements, and requires no symbols for the indices.

Abstract index notation

The abstract index notation is a way to write tensors such that the indices are no longer thought of as numerical, but rather are indeterminates. This notation captures the expressiveness of indices and the basis-independence of index-free notation.

Component-free notation

A component-free treatment of tensors uses notation that emphasises that tensors do not rely on any basis, and is defined in terms of the tensor product of vector spaces.

Operations

There are a number of basic operations that may be conducted on tensors that again produce a tensor. The linear nature of tensor implies that two tensors of the same type may be added together, and that tensors may be multiplied by a scalar with results analogous to the scaling of a vector. On components, these operations are simply performed component for component. These operations do not change the type of the tensor, however there also exist operations that change the type of the tensors.

Tensor product

The tensor product takes two tensors, S and T, and produces a new tensor, ST, whose order is the sum of the orders of the original tensors. When described as multilinear maps, the tensor product simply multiplies the two tensors, i.e.
(S\otimes T)(v_1,\ldots, v_n, v_{n+1},\ldots, v_{n+m}) = S(v_1,\ldots, v_n)T( v_{n+1},\ldots, v_{n+m}),
which again produces a map that is linear in all its arguments. On components the effect similarly is to multiply the components of the two input tensors, i.e.
(S\otimes T)^{i_1\ldots i_l i_{l+1}\ldots i_{l+n}}_{j_1\ldots j_k j_{k+1}\ldots j_{k+m}} =
S^{i_1\ldots i_l}_{j_1\ldots j_k} T^{i_{l+1}\ldots i_{l+n}}_{j_{k+1}\ldots j_{k+m}},
If S is of type (l,k) and T is of type (n,m), then the tensor product ST has type (l+n,k+m).

Contraction

Tensor contraction is an operation that reduces the total order of a tensor by two. More precisely, it reduces a type (n, m) tensor to a type (n − 1, m − 1) tensor. In terms of components, the operation is achieved by summing over one contravariant and one covariant index of tensor. For example, a (1, 1)-tensor T_i^j can be contracted to a scalar through
T_i^i.
Where the summation is again implied. When the (1, 1)-tensor is interpreted as a linear map, this operation is known as the trace.

The contraction is often used in conjunction with the tensor product to contract an index from each tensor.

The contraction can also be understood in terms of the definition of a tensor as an element of a tensor product of copies of the space V with the space V* by first decomposing the tensor into a linear combination of simple tensors, and then applying a factor from V* to a factor from V. For example, a tensor
T \in V\otimes V\otimes V^*
can be written as a linear combination
T=v_1\otimes w_1\otimes \alpha_1 + v_2\otimes w_2\otimes \alpha_2 +\cdots + v_N\otimes w_N\otimes \alpha_N.
The contraction of T on the first and last slots is then the vector
\alpha_1(v_1)w_1 + \alpha_2(v_2)w_2+\cdots+\alpha_N(v_N)w_N.

Raising or lowering an index

When a vector space is equipped with a nondegenerate bilinear form (or metric tensor as it is often called in this context), operations can be defined that convert a contravariant (upper) index into a covariant (lower) index and vice versa. A metric tensor is a (symmetric) (0, 2)-tensor, it is thus possible to contract an upper index of a tensor with one of lower indices of the metric tensor in the product. This produces a new tensor with the same index structure as the previous, but with lower index in the position of the contracted upper index. This operation is quite graphically known as lowering an index.

Conversely, the inverse operation can be defined, and is called raising an index. This is equivalent to a similar contraction on the product with a (2, 0)-tensor. This inverse metric tensor has components that are the matrix inverse of those if the metric tensor.

Applications

Continuum mechanics

Important examples are provided by continuum mechanics. The stresses inside a solid body or fluid are described by a tensor. The stress tensor and strain tensor are both second-order tensors, and are related in a general linear elastic material by a fourth-order elasticity tensor. In detail, the tensor quantifying stress in a 3-dimensional solid object has components that can be conveniently represented as a 3 × 3 array. The three faces of a cube-shaped infinitesimal volume segment of the solid are each subject to some given force. The force's vector components are also three in number. Thus, 3 × 3, or 9 components are required to describe the stress at this cube-shaped infinitesimal segment. Within the bounds of this solid is a whole mass of varying stress quantities, each requiring 9 quantities to describe. Thus, a second-order tensor is needed.

If a particular surface element inside the material is singled out, the material on one side of the surface will apply a force on the other side. In general, this force will not be orthogonal to the surface, but it will depend on the orientation of the surface in a linear manner. This is described by a tensor of type (2, 0), in linear elasticity, or more precisely by a tensor field of type (2, 0), since the stresses may vary from point to point.

Other examples from physics

Common applications include

Applications of tensors of order > 2

The concept of a tensor of order two is often conflated with that of a matrix. Tensors of higher order do however capture ideas important in science and engineering, as has been shown successively in numerous areas as they develop. This happens, for instance, in the field of computer vision, with the trifocal tensor generalizing the fundamental matrix.

The field of nonlinear optics studies the changes to material polarization density under extreme electric fields. The polarization waves generated are related to the generating electric fields through the nonlinear susceptibility tensor. If the polarization P is not linearly proportional to the electric field E, the medium is termed nonlinear. To a good approximation (for sufficiently weak fields, assuming no permanent dipole moments are present), P is given by a Taylor series in E whose coefficients are the nonlinear susceptibilities:
 \frac{P_i}{\varepsilon_0} = \sum_j  \chi^{(1)}_{ij} E_j  +  \sum_{jk} \chi_{ijk}^{(2)} E_j E_k + \sum_{jk\ell} \chi_{ijk\ell}^{(3)} E_j E_k E_\ell  + \cdots. \!
Here \chi^{(1)} is the linear susceptibility, \chi^{(2)} gives the Pockels effect and second harmonic generation, and \chi^{(3)} gives the Kerr effect. This expansion shows the way higher-order tensors arise naturally in the subject matter.

Generalizations

Tensor products of vector spaces

The vector spaces of a tensor product need not be the same, and sometimes the elements of such a more general tensor product are called "tensors". For example, an element of the tensor product space VW is a second-order "tensor" in this more general sense,[10] and an order-d tensor may likewise be defined as an element of a tensor product of d different vector spaces.[11] A type (n, m) tensor, in the sense defined previously, is also a tensor of order n + m in this more general sense.

Tensors in infinite dimensions

The notion of a tensor can be generalized in a variety of ways to infinite dimensions. One, for instance, is via the tensor product of Hilbert spaces.[12] Another way of generalizing the idea of tensor, common in nonlinear analysis, is via the multilinear maps definition where instead of using finite-dimensional vector spaces and their algebraic duals, one uses infinite-dimensional Banach spaces and their continuous dual.[13] Tensors thus live naturally on Banach manifolds.[14]

Tensor densities

The concept of a tensor field can be generalized by considering objects that transform differently. An object that transforms as an ordinary tensor field under coordinate transformations, except that it is also multiplied by the determinant of the Jacobian of the inverse coordinate transformation to the w^{\text{th}} power, is called a tensor density with weight w.[15] Invariantly, in the language of multilinear algebra, one can think of tensor densities as multilinear maps taking their values in a density bundle such as the (1-dimensional) space of n-forms (where n is the dimension of the space), as opposed to taking their values in just R. Higher "weights" then just correspond to taking additional tensor products with this space in the range.
A special case are the scalar densities. Scalar 1-densities are especially important because it makes sense to define their integral over a manifold. They appear, for instance, in the Einstein–Hilbert action in general relativity. The most common example of a scalar 1-density is the volume element, which in the presence of a metric tensor g is the square root of its determinant in coordinates, denoted \sqrt{\det g}. The metric tensor is a covariant tensor of order 2, and so its determinant scales by the square of the coordinate transition:
\det(g') = \left(\det\frac{\partial x}{\partial x'}\right)^2\det(g)
which is the transformation law for a scalar density of weight +2.

More generally, any tensor density is the product of an ordinary tensor with a scalar density of the appropriate weight. In the language of vector bundles, the determinant bundle of the tangent bundle is a line bundle that can be used to 'twist' other bundles w times. While locally the more general transformation law can indeed be used to recognise these tensors, there is a global question that arises, reflecting that in the transformation law one may write either the Jacobian determinant, or its absolute value. Non-integral powers of the (positive) transition functions of the bundle of densities make sense, so that the weight of a density, in that sense, is not restricted to integer values. Restricting to changes of coordinates with positive Jacobian determinant is possible on orientable manifolds, because there is a consistent global way to eliminate the minus signs; but otherwise the line bundle of densities and the line bundle of n-forms are distinct. For more on the intrinsic meaning, see density on a manifold.

Spinors

When changing from one orthonormal basis (called a frame) to another by a rotation, the components of a tensor transform by that same rotation. This transformation does not depend on the path taken through the space of frames. However, the space of frames is not simply connected (see orientation entanglement and plate trick): there are continuous paths in the space of frames with the same beginning and ending configurations that are not deformable one into the other. It is possible to attach an additional discrete invariant to each frame called the "spin" that incorporates this path dependence, and which turns out to have values of ±1. A spinor is an object that transforms like a tensor under rotations in the frame, apart from a possible sign that is determined by the spin.

History

The concepts of later tensor analysis arose from the work of Carl Friedrich Gauss in differential geometry, and the formulation was much influenced by the theory of algebraic forms and invariants developed during the middle of the nineteenth century.[16] The word "tensor" itself was introduced in 1846 by William Rowan Hamilton[17] to describe something different from what is now meant by a tensor.[Note 6] The contemporary usage was introduced by Woldemar Voigt in 1898.[18]

Tensor calculus was developed around 1890 by Gregorio Ricci-Curbastro under the title absolute differential calculus, and originally presented by Ricci in 1892.[19] It was made accessible to many mathematicians by the publication of Ricci and Tullio Levi-Civita's 1900 classic text Méthodes de calcul différentiel absolu et leurs applications (Methods of absolute differential calculus and their applications).[20]

In the 20th century, the subject came to be known as tensor analysis, and achieved broader acceptance with the introduction of Einstein's theory of general relativity, around 1915. General relativity is formulated completely in the language of tensors. Einstein had learned about them, with great difficulty, from the geometer Marcel Grossmann.[21] Levi-Civita then initiated a correspondence with Einstein to correct mistakes Einstein had made in his use of tensor analysis. The correspondence lasted 1915–17, and was characterized by mutual respect:
I admire the elegance of your method of computation; it must be nice to ride through these fields upon the horse of true mathematics while the like of us have to make our way laboriously on foot.
—Albert Einstein, The Italian Mathematicians of Relativity[22]
Tensors were also found to be useful in other fields such as continuum mechanics. Some well-known examples of tensors in differential geometry are quadratic forms such as metric tensors, and the Riemann curvature tensor. The exterior algebra of Hermann Grassmann, from the middle of the nineteenth century, is itself a tensor theory, and highly geometric, but it was some time before it was seen, with the theory of differential forms, as naturally unified with tensor calculus. The work of Élie Cartan made differential forms one of the basic kinds of tensors used in mathematics.

From about the 1920s onwards, it was realised that tensors play a basic role in algebraic topology (for example in the Künneth theorem).[23] Correspondingly there are types of tensors at work in many branches of abstract algebra, particularly in homological algebra and representation theory. Multilinear algebra can be developed in greater generality than for scalars coming from a field. For example, scalars can come from a ring. But the theory is then less geometric and computations more technical and less algorithmic.[24] Tensors are generalized within category theory by means of the concept of monoidal category, from the 1960s.[25]

Australia’s scorching 2013 heat record was ‘virtually impossible’ without global warming

February 9
Original link:  http://www.washingtonpost.com/news/energy-environment/wp/2015/02/09/australias-scorching-2013-heat-record-was-virtually-impossible-without-global-warming/
 


Afterwards — after the blistering heat and the bushfires – they would call it the “Angry Summer.”
 
But whatever the description, the Australian summer of 2012-2013 provided a terrifying preview of a world under climate change. According to the country’s Bureau of Meteorology, a devastating heat wave in late 2012-early 2013 saw “records set in every State and Territory … the nationally averaged daily temperature rose to levels never previously observed, and did this for an extended period.”

What was truly astonishing is just how extreme the records were that summer — 44 different locations set all time high temperature records, including Sydney — and how long the heat lasted. “January 2013 brought record-setting heat to Australia; not just for days, but for weeks,” noted NASA.

The scorching temperatures at the start of 2013 then helped pave the way for a new national temperature record in 2013, surpassing the previous Australian heat record set just a few years earlier in 2005.

Here’s a NASA image of Australian temperature anomalies from Jan. 1-8, 2013:

And now, says a new report, it’s pretty hard to see how it could have all happened unless climate change was stacking the odds. The report, by the country’s independent Climate Council, finds that 2013, Australia’s record hottest year ever, was “virtually impossible” without climate change. And as for the 2012/2013 summer heat waves? Global warming made their frequency three times as likely, and doubled the chances that they’d reach the extreme heat intensity that they did.

The report was authored by Will Steffen, an adjunct professor who studies Earth system science at the Australian National University.*

So how can Steffen know how global warming changed the odds that the summer of 2012-2013′s “exceptionally large number of record high temperatures,” as he puts it, would occur?

The answer is that researchers have run many, many climate change models — high-powered simulations that use our understanding of the physics of the atmosphere and oceans — which can simulate temperatures with and without human-added climate factors. This allows them to compare just how often a given set of temperatures are expected to occur in the greenhouse enhanced world, versus a world in which somehow, we didn’t pump all that stuff up into the atmosphere.

This led to the following results:
* In a world without global warming, 2013′s record Australian temperature would only happen once out of every 12,300 years.
* The same approach suggests that global warming upped the odds of the Angry Summer’s heat waves by a factor of 3 or 2, depending on whether you’re referring to the frequency of heat waves or the magnitude of the heat waves themselves.
This week, Australia’s prime minister, Tony Abbott – who actually reversed the country’s carbon taxsurvived a leadership challenge from his own party.

* Correction: This post previously described Will Steffen as the director of the Climate Change Institute at the Australian National University. He left that post in 2012.

Chris Mooney reports on science and the environment.

Analytical skill

From Wikipedia, the free encyclopedia https://en.wikipedia.org/wiki/Analytical_skill ...