Slutsky equation

Last updated

In microeconomics, the Slutsky equation (or Slutsky identity), named after Eugen Slutsky, relates changes in Marshallian (uncompensated) demand to changes in Hicksian (compensated) demand, which is known as such since it compensates to maintain a fixed level of utility.

Contents

There are two parts of the Slutsky equation, namely the substitution effect, and income effect. In general, the substitution effect can be negative for consumers as it can limit choices. He designed this formula to explore a consumer's response as the price changes. When the price increases, the budget set moves inward, which also causes the quantity demanded to decrease. In contrast, when the price decreases, the budget set moves outward, which leads to an increase in the quantity demanded. The substitution effect is due to the effect of the relative price change while the income effect is due to the effect of income being freed up. The equation demonstrates that the change in the demand for a good, caused by a price change, is the result of two effects:

The Slutsky equation decomposes the change in demand for good i in response to a change in the price of good j:

where is the Hicksian demand and is the Marshallian demand, at the vector of price levels , wealth level (or, alternatively, income level) , and fixed utility level given by maximizing utility at the original price and income, formally given by the indirect utility function . The right-hand side of the equation is equal to the change in demand for good i holding utility fixed at u minus the quantity of good j demanded, multiplied by the change in demand for good i when wealth changes.

The first term on the right-hand side represents the substitution effect, and the second term represents the income effect. [1] Note that since utility is not observable, the substitution effect is not directly observable, but it can be calculated by reference to the other two terms in the Slutsky equation, which are observable. This process is sometimes known as the Hicks decomposition of a demand change. [2]

The equation can be rewritten in terms of elasticity:

where εp is the (uncompensated) price elasticity, εph is the compensated price elasticity, εw,i the income elasticity of good i, and bj the budget share of good j.

Overall, in simple words, the Slutsky equation states the total change in demand consists of an income effect and a substitution effect and both effects collectively must equal the total change in demand.

The equation above is helpful as it represents the fluctuation in demand are indicative of different types of good. The substitution effect will always turn out negative as indifference curves are always downward sloping. However, the same does not apply to income effect as it depends on how consumption of a good changes with income.

The income effect on a normal goods is negative, and if the price decreases, consequently purchasing power or income goes up. The reverse holds when price increases and purchasing power or income decreases, as a result of, so does demand.

Generally, not all goods are "normal". While in an economic sense, some are inferior. However, that does not equate quality-wise that they are poor rather that it sets a negative income profile - as income increases, consumers consumption of the good decreases.

For example, consumers who are running low of money for food purchase instant noodles, however, the product is not generally held as something people would normally consume on a daily basis. This is due to the constrains in terms of money; as wealth increases, consumption decreases. In this case, the substitution effect is negative, but the income effect is also negative.

In any case the substitution effect or income effect are positive or negative when prices increase depends on the type of goods:

Total EffectSubstitution EffectIncome Effect
+ Substitute goods Substitute goods Inferior goods
- Complementary goods Complementary goods Normal goods

However, whether the total effect will always be negative is impossible to tell if inferior complementary goods are mentioned. For instance, the substitution effect and the income effect pull in opposite directions. The total effect will depend on which effect is ultimately stronger.

Derivation

While there are several ways to derive the Slutsky equation, the following method is likely the simplest. Begin by noting the identity where is the expenditure function, and u is the utility obtained by maximizing utility given p and w. Totally differentiating with respect to pj yields as the following:

.

Making use of the fact that by Shephard's lemma and that at optimum,

where is the indirect utility function,

one can substitute and rewrite the derivation above as the Slutsky equation.

The Slutsky matrix

The Slutsky equation can be rewritten in matrix form:

where Dp is the derivative operator with respect to prices and Dw is the derivative operator with respect to wealth.

The matrix is known as the Hicksian substitution matrix and is formally defined as:

The Slutsky matrix is given by:

When is the maximum utility the consumer achieves at prices and income , that is, , the Slutsky equation implies that each element of the Slutsky matrix is exactly equal to the corresponding element of the Hicksian substitution matrix . The Slutsky matrix is symmetric, and given that the expenditure function is concave, the Slutsky matrix is also negative semi-definite.

Example

A Cobb-Douglas utility function (see Cobb-Douglas production function) with two goods and income generates Marshallian demand for goods 1 and 2 of and Rearrange the Slutsky equation to put the Hicksian derivative on the left-hand-side yields the substitution effect:

Going back to the original Slutsky equation shows how the substitution and income effects add up to give the total effect of the price rise on quantity demanded:

Thus, of the total decline of in quantity demanded when rises, 21/70 is from the substitution effect and 49/70 from the income effect. Good 1 is the good this consumer spends most of his income on (), which is why the income effect is so large.

One can check that the answer from the Slutsky equation is the same as from directly differentiating the Hicksian demand function, which here is [3]

where is utility. The derivative is

so since the Cobb-Douglas indirect utility function is and when the consumer uses the specified demand functions, the derivative is:

which is indeed the Slutsky equation's answer.

The Slutsky equation also can be applied to compute the cross-price substitution effect. One might think it was zero here because when rises, the Marshallian quantity demanded of good 1, is unaffected (), but that is wrong. Again rearranging the Slutsky equation, the cross-price substitution effect is:

This says that when rises, there is a substitution effect of towards good 1. At the same time, the rise in has a negative income effect on good 1's demand, an opposite effect of the exact same size as the substitution effect, so the net effect is zero. This is a special property of the Cobb-Douglas function.

Changes in multiple prices at once

When there are two goods, the Slutsky equation in matrix form is: [4]

Although strictly speaking the Slutsky equation only applies to infinitesimal changes in prices, it is standardly used a linear approximation for finite changes. If the prices of the two goods change by and , the effect on the demands for the two goods are:

Multiplying out the matrices, the effect on good 1, for example, would be

The first term is the substitution effect. The second term is the income effect, composed of the consumer's response to income loss times the size of the income loss from each price's increase.

Giffen goods

A Giffen good is a product that is in greater demand when the price increases, which are also special cases of inferior goods. [5] In the extreme case of income inferiority, the size of income effect overpowers the size of the substitution effect, leading to a positive overall change in demand responding to an increase in the price. Slutsky's decomposition of the change in demand into a pure substitution effect and income effect explains why the law of demand doesn't hold for Giffen goods.

See also

Related Research Articles

In mathematics, the Laplace operator or Laplacian is a differential operator given by the divergence of the gradient of a scalar function on Euclidean space. It is usually denoted by the symbols , (where is the nabla operator), or . In a Cartesian coordinate system, the Laplacian is given by the sum of second partial derivatives of the function with respect to each independent variable. In other coordinate systems, such as cylindrical and spherical coordinates, the Laplacian also has a useful form. Informally, the Laplacian Δf (p) of a function f at a point p measures by how much the average value of f over small spheres or balls centered at p deviates from f (p).

In continuum mechanics, the infinitesimal strain theory is a mathematical approach to the description of the deformation of a solid body in which the displacements of the material particles are assumed to be much smaller than any relevant dimension of the body; so that its geometry and the constitutive properties of the material at each point of space can be assumed to be unchanged by the deformation.

A good's price elasticity of demand is a measure of how sensitive the quantity demanded is to its price. When the price rises, quantity demanded falls for almost any good, but it falls more for some than for others. The price elasticity gives the percentage change in quantity demanded when there is a one percent increase in price, holding everything else constant. If the elasticity is −2, that means a one percent price rise leads to a two percent decline in quantity demanded. Other elasticities measure how the quantity demanded changes with other variables.

<span class="mw-page-title-main">Hooke's law</span> Physical law: force needed to deform a spring scales linearly with distance

In physics, Hooke's law is an empirical law which states that the force needed to extend or compress a spring by some distance scales linearly with respect to that distance—that is, Fs = kx, where k is a constant factor characteristic of the spring, and x is small compared to the total possible deformation of the spring. The law is named after 17th-century British physicist Robert Hooke. He first stated the law in 1676 as a Latin anagram. He published the solution of his anagram in 1678 as: ut tensio, sic vis. Hooke states in the 1678 work that he was aware of the law since 1660.

Linear elasticity is a mathematical model of how solid objects deform and become internally stressed due to prescribed loading conditions. It is a simplification of the more general nonlinear theory of elasticity and a branch of continuum mechanics.

<span class="mw-page-title-main">Euler equations (fluid dynamics)</span> Set of quasilinear hyperbolic equations governing adiabatic and inviscid flow

In fluid dynamics, the Euler equations are a set of quasilinear partial differential equations governing adiabatic and inviscid flow. They are named after Leonhard Euler. In particular, they correspond to the Navier–Stokes equations with zero viscosity and zero thermal conductivity.

In Hamiltonian mechanics, a canonical transformation is a change of canonical coordinates that preserves the form of Hamilton's equations. This is sometimes known as form invariance. It need not preserve the explicit form of the written-out Hamiltonian itself, just the form of the equations in which it appears as a parameter. Canonical transformations are useful in their own right, and also form the basis for the Hamilton–Jacobi equations and Liouville's theorem.

<span class="mw-page-title-main">Law of demand</span> Fundamental principle in microeconomics

In microeconomics, the law of demand is a fundamental principle which states that there is an inverse relationship between price and quantity demanded. In other words, "conditional on all else being equal, as the price of a good increases (↑), quantity demanded will decrease (↓); conversely, as the price of a good decreases (↓), quantity demanded will increase (↑)". Alfred Marshall worded this as: "When we say that a person's demand for anything increases, we mean that he will buy more of it than he would before at the same price, and that he will buy as much of it as before at a higher price". The law of demand, however, only makes a qualitative statement in the sense that it describes the direction of change in the amount of quantity demanded but not the magnitude of change.

In physics, the Hamilton–Jacobi equation, named after William Rowan Hamilton and Carl Gustav Jacob Jacobi, is an alternative formulation of classical mechanics, equivalent to other formulations such as Newton's laws of motion, Lagrangian mechanics and Hamiltonian mechanics.

<span class="mw-page-title-main">Total least squares</span>

In applied statistics, total least squares is a type of errors-in-variables regression, a least squares data modeling technique in which observational errors on both dependent and independent variables are taken into account. It is a generalization of Deming regression and also of orthogonal regression, and can be applied to both linear and non-linear models.

In microeconomics, a consumer's Marshallian demand function is the quantity they demand of a particular good as a function of its price, their income, and the prices of other goods, a more technical exposition of the standard demand function. It is a solution to the utility maximization problem of how the consumer can maximize their utility for given income and prices. A synonymous term is uncompensated demand function, because when the price rises the consumer is not compensated with higher nominal income for the fall in their real income, unlike in the Hicksian demand function. Thus the change in quantity demanded is a combination of a substitution effect and a wealth effect. Although Marshallian demand is in the context of partial equilibrium theory, it is sometimes called Walrasian demand as used in general equilibrium theory.

In mechanics, virtual work arises in the application of the principle of least action to the study of forces and movement of a mechanical system. The work of a force acting on a particle as it moves along a displacement is different for different displacements. Among all the possible displacements that a particle may follow, called virtual displacements, one will minimize the action. This displacement is therefore the displacement followed by the particle according to the principle of least action.

The work of a force on a particle along a virtual displacement is known as the virtual work.

In numerical analysis, the Crank–Nicolson method is a finite difference method used for numerically solving the heat equation and similar partial differential equations. It is a second-order method in time. It is implicit in time, can be written as an implicit Runge–Kutta method, and it is numerically stable. The method was developed by John Crank and Phyllis Nicolson in the mid 20th century.

In microeconomics, a consumer's Hicksian demand function or compensated demand function for a good is his quantity demanded as part of the solution to minimizing his expenditure on all goods while delivering a fixed level of utility. Essentially, a Hicksian demand function shows how an economic agent would react to the change in the price of a good, if the agent's income was compensated to guarantee the agent the same utility previous to the change in the price of the good—the agent will remain on the same indifference curve before and after the change in the price of the good. The function is named after John Hicks.

Roy's identity is a major result in microeconomics having applications in consumer choice and the theory of the firm. The lemma relates the ordinary (Marshallian) demand function to the derivatives of the indirect utility function. Specifically, denoting the indirect utility function as the Marshallian demand function for good can be calculated as

In mathematics, the discrete Poisson equation is the finite difference analog of the Poisson equation. In it, the discrete Laplace operator takes the place of the Laplace operator. The discrete Poisson equation is frequently used in numerical analysis as a stand-in for the continuous Poisson equation, although it is also studied in its own right as a topic in discrete mathematics.

The Rabi problem concerns the response of an atom to an applied harmonic electric field, with an applied frequency very close to the atom's natural frequency. It provides a simple and generally solvable example of light–atom interactions and is named after Isidor Isaac Rabi.

Non-linear least squares is the form of least squares analysis used to fit a set of m observations with a model that is non-linear in n unknown parameters (m ≥ n). It is used in some forms of nonlinear regression. The basis of the method is to approximate the model by a linear one and to refine the parameters by successive iterations. There are many similarities to linear least squares, but also some significant differences. In economic theory, the non-linear least squares method is applied in (i) the probit regression, (ii) threshold regression, (iii) smooth regression, (iv) logistic link regression, (v) Box–Cox transformed regressors ().

<span class="mw-page-title-main">Interval finite element</span>

In numerical analysis, the interval finite element method is a finite element method that uses interval parameters. Interval FEM can be applied in situations where it is not possible to get reliable probabilistic characteristics of the structure. This is important in concrete structures, wood structures, geomechanics, composite structures, biomechanics and in many other areas. The goal of the Interval Finite Element is to find upper and lower bounds of different characteristics of the model and use these results in the design process. This is so called worst case design, which is closely related to the limit state design.

<span class="mw-page-title-main">Derivations of the Lorentz transformations</span>

There are many ways to derive the Lorentz transformations using a variety of physical principles, ranging from Maxwell's equations to Einstein's postulates of special relativity, and mathematical tools, spanning from elementary algebra and hyperbolic functions, to linear algebra and group theory.

References

  1. Nicholson, W. (2005). Microeconomic Theory (10th ed.). Mason, Ohio: Thomson Higher Education.
  2. Varian, H. (1992). Microeconomic Analysis (3rd ed.). New York: W. W. Norton.
  3. Varian, H. (1992). Microeconomic Analysis (3rd ed.). New York: W. W. Norton., p. 121.
  4. Varian, H. (1992). Microeconomic Analysis (3rd ed.). New York: W. W. Norton., pp. 120-121.
  5. Varian, Hal R. “Chapter 8: Slutsky Equation.” Essay. In Intermediate Microeconomics with Calculus, 1st ed., 137. New York, NY: W W Norton, 2014.

References

Varian, H. R. (2020). Intermediate microeconomics : a modern approach (Ninth edition.). W.W. Norton & Company.