Skip to main content
Log in

Welfare evaluations and price indices with path dependency problems

  • Original Paper
  • Published:
Social Choice and Welfare Aims and scope Submit manuscript

Abstract

In cases where multiple prices change, this paper develops methods for calculating compensating and equivalent variations from the prices, the consumption bundle, and the wealth elasticities of demand for goods. The methods are a natural extension and generalization of the method in Willig (Am Econ Rev 66(4):589–597, 1976), and have the ability to provide second-order approximations of compensating and equivalent variations. In addition, this paper considers two types of price paths: indifference price paths, along which utility levels are kept constant, and iso-price-ratio paths, along which prices change proportionally. Along these two paths, changes in consumer surplus, compensating variation, and equivalent variation are easily calculated, and that there are interesting and precise relationships among them. Furthermore, the methods and relationships above correspond to those of the Laspeyres–Konüs and Paasche–Konüs true cost of living price indices and the Divisia indices.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. The concept of consumer surplus was introduced by Dupuit (1844). Hicks (1941, 1946) and Marshall (1961) advanced and analyzed the concept of consumer surplus. Harberger (1971) showed that consumer surplus has some desirable properties and that it is superior to gross national income in many ways. Hicks (1946) introduced compensating and equivalent variations.

  2. In recent years, various authors have advanced the research on consumer surplus. Blackorby and Donaldson (1999) showed that the area to the left of market demand curves in directly affected markets is equal to the sum of consumer and producer surpluses in all markets in general equilibrium settings. Vives (1987), Miyake (2006) and Hayashi (2013) examined under what conditions income effects on commodities under consideration vanish.

  3. McKenzie (1976) and McKenzie and Pearce (1976, 1982) developed the method for expressing compensating and equivalent variations as the Taylor series of the indirect utility function. Using Hurwicz and Uzawa (1971) results, Hausman (1981) derived the compensating and equivalent variations from an estimate of the demand curve.

  4. The Divisia price index is closely related to changes in consumer surplus. As Hulten (1973) pointed out, the Divisia price index has the same path dependency property as consumer surplus.

  5. Price indices are also used for comparing standards of living between countries. For example, see Deaton and Heston (2010).

  6. Diewert (1988) is a great summary of the early history of price index research. When it comes to recent developments in price index theory, see Moulton (1996), Nordhaus (1998), Pollak (1998), Persky (1998), Boskin et al. (1998), Abraham et al. (1998), Diewert (1998), Schultze (2003), Hausman (2003), Abraham (2003) and Lebow and Rudd (2003).

  7. When the utility function is homothetic, the superlative indices proposed by Diewert (1976) provide close approximations of true cost of living indices.

  8. As will be shown in Appendix C.1, the Laspeyres–Konüs and Paasche–Konüs true cost of living indices are closely related to compensating and equivalent variations, and the Divisia price index is closely related to changes in consumer surplus.

  9. In this paragraph, I focus on compensating variation, equivalent variation, and consumer surplus. However, similar local and global properties about price indices also hold true because there are close relationships between these three measures of economic welfare and the corresponding price and quantity indices, as explained in Appendix C.

  10. We regard vectors as column vectors.

  11. When there is no money illusion, without loss of generality, we can normalize the income level to one by replacing p with \(p / ( p\cdot x )\). When the consumer’s income level changes, we can adjust the prices for the changes.

  12. When income level w is normalized to one, and there is no need to express the Marshallian demand function as x(pw) and the inverse demand function as p(xw), we denote the Marshallian demand function by x(p) and the inverse demand function by p(x), respectively. Since the consumer has a rational, strongly monotone, locally nonsatiated, and strictly convex preference relation \(\succsim \), x(p) and p(x) are bijections.

  13. In general, we use \(C_{p}(i, p^{1}, p^{2})\) and \(C_{p}(i)\) as both a curve and a point on a curve interchangeably. \(C_{p}\) is used as a curve: \(C_{p}=\{ p\in \mathbb {R}_{++}^{N}| p=C_{p}(i; p^{1}, p^{2}), s\le i \le t\}\).

  14. The Divisia price index, which will be explained later, also has a path dependency problem.

  15. As mentioned before, compensating and equivalent variations are path independent.

  16. The approach in Willig (1976) and mine differ a bit. Willig approximated the lower and upper bounds of compensating and equivalent variations with changes in consumer surplus, using the lower and upper bounds of the wealth elasticity of demand. On the other hand, given a price change, I calculate the approximations of compensating and equivalent variations.

  17. \(C_{x}^{j}\) in the following equations is defined from \(C_{x}^{j}(i; x^{1}, x^{2})\) in the same way.

  18. In Appendix A, using manifolds and differential forms, the differences among the changes in consumer surplus along different paths are analyzed in a more general way. In Appendix A.4, the proof of Proposition 7 is given.

  19. Of course, some of the following equations in this and following subsections are approximations. Thus, there would be slightly different ways to express these equations.

  20. The corresponding Laspeyres–Konüs and Paasche–Konüs true cost of living price indices are stated in Appendix C.2.

  21. In Eqs. (10), (11), (16), and (17), we ignore small terms that contain three or more \(\varDelta p\) or \(\varDelta x\), like \((\varDelta p \cdot x^{1, 1})(\varDelta p \cdot \varDelta x)\).

  22. In response to McKenzie (1979), Willig (1979) also provided the following methodology for calculating the upper and lower bounds on the errors of the approximations of compensating variations with changes in consumer surplus when multiple prices change:

    $$\begin{aligned}&C-A \ge \frac{\underline{\mu }_{u}A^{2}_{u}}{2m^{0}}+\frac{\underline{\mu }_{d}A^{2}_{d}}{2m^{0}}+\frac{\overline{\mu }_{d}A_{u}A_{d}}{m^{0}}\left( 1+\frac{\overline{\mu }_{u}A_{u}}{2m^{0}} \right) +0.01A_{d}-0.005A_{u} \end{aligned}$$
    (18)
    $$\begin{aligned}&C-A \le \frac{\overline{\mu }_{u}A^{2}_{u}}{2m^{0}}+\frac{\overline{\mu }_{d}A^{2}_{d}}{2m^{0}}+\frac{\underline{\mu }_{d}A_{u}A_{d}}{m^{0}}\left( 1+\frac{\underline{\mu }_{u}A_{u}}{2m^{0}} \right) -0.015A_{d}+0.005A_{u}, \end{aligned}$$
    (19)

    where \(A_{i}, i=u, d\) is the change in consumer surplus caused by price changes; \(\overline{\mu }_{i}, i=u, d\) and \(\underline{\mu }_{i}, i=u, d\) are the upper and lower bounds of the income elasticity of demand; and the subscripts u and d stand for “prices up” and “prices down.” In this estimation, Willig divided all goods into the “prices up” and the “prices down” categories and estimated the upper and lower bounds of the errors.

    However, in comparison to Eqs. (16) and (17), Eqs. (18) and (19) have flaws. First, since Willig divided all goods into two types, the error margins of the approximations might be very large when more than two prices change. Second, the approximation of the utility function causes unnecessary errors, such as \(0.01A_{d}\) and \(0.005A_{u}\). Substituting zero into \(A_{u}\) or \(A_{d}\) in Eqs. (18) and (19), it is clear that (18) and (19) are not consistent with (9).

  23. A more rigorous definition of consumption surface is given in Appendix A.1.

  24. If \(\varDelta x^{i, i}\) is small enough, the area enclosed by four points \((x^{i, i}, x^{i+1, i}, x^{i, i+1}, x^{i+1, i+1})\) is assumed to be a parallelogram. It is assumed that \(\varDelta x^{i, i}, \forall i\) divides the parallelogram with four points \((x^{i, i}, x^{i+1, i}, x^{i, i+1}, x^{i+1, i+1})\) into two parts whose areas are the same.

  25. The following equation corresponds to Corollary 2 in Appendix A.2.

  26. In a more general form, the following equations also hold true:

    $$\begin{aligned}&\log ( r(p^{k+n, l}, p^{k+n, l+m}) )-\log ( r(p^{k, l}, p^{k, l+m}) ) =-\sum _{j=1}^{m}\sum _{i=1}^{n}\psi ^{k+i-1, l+j-1} \\&\quad \frac{r(p^{k+n, l}, p^{k+n, l+m}) }{ r(p^{k, l}, p^{k, l+m})}=\exp \left( -\sum _{j=1}^{m}\sum _{i=1}^{n}\psi ^{k+i-1, l+j-1} \right) . \end{aligned}$$
  27. See Appendix B.

  28. The corresponding Laspeyres–Konüs and Paasche–Konüs true cost of living price indices are stated in Appendix C.2.

  29. Since it is apparent that if a property about consumer surplus, compensating variation, or equivalent variation holds true, a similar property about the corresponding price and quantity indices holds true from the discussion in Appendix C.1. Hereafter in this subsection, I focus on the properties of consumer surplus, compensating variations, and equivalent variations and their calculation methods.

  30. See Appendix A.

  31. For the same reason, this paper’s methods for calculating true cost of living indices are able to give superior estimates to Diewert’s superlative indices, even though they need more information about the demand and inverse demand function than Diewert’s, as explained below.

  32. It is also important to note that since this example has only one parameter, \(\theta \), the specification is really restrictive; thus, unless you are confident about the functional form of the underlying utility function, this CES functional form might cause large misestimation of compensating and equivalent variations regardless of whether price changes are small or large.

  33. As Proposition 4 shows, when the prices change along an iso-price-ratio path, the change in consumer surplus, the compensating variation, and the equivalent variation are easily convertible to each other. And as Proposition 1 implies, the change in consumer surplus along an indifference price path is zero. Thus, these definitions make it easier to compare the change in consumer surplus, the compensating variation, and the equivalent variation when the prices change from \(p^{1, 1}\) to \(p^{2, 2}\).

  34. As the definition of \(S^{x}(p^{1, 1}, p^{2, 2})\) implies, there could be infinitely many smooth consumption surfaces generated by the price change from \(p^{1, 1}\) to \(p^{2, 2}\). Thus, we can choose any one of them. However, if the prices change along a price path, the smooth consumption surface that contains the price path should be chosen so that the relationships among the change in consumer surplus, the compensating variation, and the equivalent variation are easily analyzed.

  35. x in \(x_{w}(p(x), w)\) will be omitted when unnecessary.

  36. We use the word ‘submanifold’ because a smooth consumption surface could be considered as a subset of the consumption space, which is a manifold.

  37. The equation is the same as Eq. (1) so that the equation calculates changes in consumer surplus along consumption paths.

  38. We choose the sign of Eq. (A-3) such that it is consistent with the sign of Eq. (A-4).

  39. As I have said, we assume that the income level is normalized to one: \(w=1\).

  40. Equation (A-6) implies \(p (x, w) \parallel q^{k} (x, w)\) for \(k = 1, 2\), where \(q^{k}(x, w)\) for \(k= 1, 2\) is an N-dimensional vector and defined by

    $$\begin{aligned} q_{i}^{k} (x, w) = \sum _{j = 1}^{N} \left( \frac{\partial p_{i} (x, w)}{\partial x_{j}} - \frac{\partial p_{j} (x, w)}{\partial x_{i}} \right) c_{j}^{k},\quad k=1, 2. \end{aligned}$$
  41. Since we consider infinitely small change cases now, we assume that \(p^{1, 1}\cdot \varDelta x=- \varDelta p \cdot x^{1, 1}\).

  42. As M increases, \(( 1+ \alpha )^{(1 / z)}-1\) approaches \((1 / M)\log (1+\alpha )\).

  43. Since the income level along the price path is always one, \(\sum _{j=1}^{N} x_{j}(q(i), w )q_{j}(i)=1, \forall i\) holds true.

References

  • Abraham KG (2003) Toward a cost-of-living index: progress and prospects. J Econ Perspect 17(1):45–58

    Article  Google Scholar 

  • Abraham KG, Greenlees JS, Moulton BR (1998) Working to improve the consumer price index. J Econ Perspect 12(1):27–36

    Article  Google Scholar 

  • Blackorby C, Donaldson D (1999) Market demand curves and Dupuit–Marshall consumers’ surpluses: a general equilibrium analysis. Math Soc Sci 37(2):139–163

    Article  Google Scholar 

  • Boskin MJ, Dulberger ER, Gordon RJ, Griliches Z, Jorgenson DW (1998) Consumer prices, the consumer price index, and the cost of living. J Econ Perspect 12(1):3–26

    Article  Google Scholar 

  • Bruce N (1977) A note on consumer’s surplus, the divisia index, and the measurement of welfare changes. Econometr J Econometr Soc 45(4):1033–1038

    Article  Google Scholar 

  • Chipman JS, Moore JC (1976) The scope of consumer’s surplus arguments. In: Tang AM, Westfield FM, Worley JS (eds) Welfare and time in economics: essays in honor of Nicholas Georgescu-Roegen. Lexington Books, Lexington, pp 69–123

    Google Scholar 

  • Deaton A, Heston A (2010) Understanding PPPs and PPP-based national accounts. Am Econ J Macroecon 2(4):1–35

    Article  Google Scholar 

  • Diewert WE (1976) Exact and superlative index numbers. J Econometr 4(2):115–145

    Article  Google Scholar 

  • Diewert WE (1988) The early history of price index research. https://doi.org/10.3386/w2713

  • Diewert WE (1998) Index number issues in the consumer price index. J Econ Perspect 12(1):47–58

    Article  Google Scholar 

  • Dupuit J (1844) On the measurement of the utility of public works. Int Econ Pap 2:83–110 [translated and reprinted in Readings in Welfare Economics (Kenneth J. Arrow and Tibor Scitovsky, eds.), 1969, 255–283, Homewood]

  • Harberger AC (1971) Three basic postulates for applied welfare economics: an interpretive essay. J Econ Lit 9(3):785–797

    Google Scholar 

  • Hausman JA (1981) Exact consumer’s surplus and deadweight loss. Am Econ Rev 71(4):662–676

    Google Scholar 

  • Hausman JA (2003) Sources of bias and solutions to bias in the consumer price index. J Econ Perspect 17(1):23–44

    Article  Google Scholar 

  • Hayashi T (2013) Smallness of a commodity and partial equilibrium analysis. J Econ Theory 148(1):279–305

    Article  Google Scholar 

  • Hicks JR (1941) The rehabilitation of consumers’ surplus. Rev Econ Stud 8(2):108–116

    Article  Google Scholar 

  • Hicks JR (1946) Value and capital: an inquiry into some fundamental principles of economic theory. Clarendon Press, Oxford

    Google Scholar 

  • Hulten CR (1973) Divisia index numbers. Econometr J Econometr Soc 41(6):1017–1025

    Article  Google Scholar 

  • Hurwicz L, Uzawa H (1971) On the integrability of demand functions. In: Chipman JS (ed) Preferences, utility, and demand. Harcourt Brace Jovanovich, New York, pp 114–148

    Google Scholar 

  • Lebow DE, Rudd JB (2003) Measurement error in the consumer price index: where do we stand? J Econ Lit 41(1):159–201

    Article  Google Scholar 

  • Marshall A (1961) Principles of economics, 9th edn. Macmillan, New York (first published in 1890)

    Google Scholar 

  • Mas-Colell A, Whinston MD, Green JR (1995) Microeconomic theory, vol 1. Oxford University Press, New York

    Google Scholar 

  • McKenzie GW (1976) Measuring gains and losses. J Polit Econ 84(3):641–646

    Article  Google Scholar 

  • McKenzie GW (1979) Consumer’s surplus without apology: comment. Am Econ Rev 69(3):465–468

    Google Scholar 

  • McKenzie GW, Pearce IF (1976) Exact measures of welfare and the cost of living. Rev Econ Stud 43(3):465–468

    Article  Google Scholar 

  • McKenzie GW, Pearce IF (1982) Welfare measurement: a synthesis. Am Econ Rev 72(4):669–682

    Google Scholar 

  • Miyake M (2006) On the applicability of Marshallian partial-equilibrium analysis. Math Soc Sci 52(2):176–196

    Article  Google Scholar 

  • Moulton BR (1996) Bias in the consumer price index: what is the evidence? J Econ Perspect 10(4):159–177

    Article  Google Scholar 

  • Nordhaus WD (1998) Quality change in price indexes. J Econ Perspect 12(1):59–68

    Article  Google Scholar 

  • Persky J (1998) Retrospectives: price indexes and general exchange values. J Econ Perspect 12(1):197–205

    Article  Google Scholar 

  • Pollak RA (1998) The consumer price index: a research agenda and three proposals. J Econ Perspect 12(1):69–78

    Article  Google Scholar 

  • Schultze CL (2003) The consumer price index: conceptual issues and practical suggestions. J Econ Perspect 17(1):3–22

    Article  Google Scholar 

  • Silberberg E (1972) Duality and the many consumer’s surpluses. Am Econ Rev 62(5):942–952

    Google Scholar 

  • Vives X (1987) Small income effects: a Marshallian theory of consumer surplus and downward sloping demand. Rev Econ Stud 54(1):87–103

    Article  Google Scholar 

  • Willig RD (1976) Consumer’s surplus without apology. Am Econ Rev 66(4):589–597

    Google Scholar 

  • Willig RD (1979) Consumer’s surplus without apology: reply. Am Econ Rev 69(3):469–474

    Google Scholar 

Download references

Acknowledgements

I am deeply grateful to Robert D. Willig, Makoto Mori, Tetsuya Nakajima, Minoru Kitahara, Ryosuke Okazawa, the editors, and anonymous reviewers for their helpful comments and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tsuyoshi Sasaki.

Appendices

Appendix A: Proofs and explanations using manifold and differential forms

1.1 A.1 Definitions and Stokes’ theorem

Let M be a two dimensional smooth submanifold with boundary that is embedded in an N-dimensional Euclidean space \(\mathbb {R}^{N}\), and let \(\partial M\) be the boundary of the submanifold M. Define \(T_{x}M\) as the tangent space of M at x, and define \(T^{*}_{x}M\) as the cotangent space of M at x. Let \(TM=\bigcup _{x \in M} T_{x}M\) be the tangent bundle, and let \(T^{*}M=\bigcup _{x \in M}T^{*}_{x}M\) be the cotangent bundle.

Here, well known basic results on manifold and differential forms are explained and described. Let \(\omega :M \rightarrow T^{*}M\) be a 1-form on M. Let \(x_{1}, x_{2}, x_{3}, \ldots , x_{N}\) be the standard coordinates on \(\mathbb {R}^{N}\). At each point \(x\in M\), the value of the 1-form \(\omega \) can be written as a linear combination:

$$\begin{aligned} \omega _{x}= \sum _{i=1}^{N}a_{i}(x)dx_{i}, \end{aligned}$$

where \(dx_{1}, dx_{2}, dx_{3}, \ldots , dx_{N}\) are the basis for the cotangent space \(T^{*}_{x}M\). The exterior derivative of \(\omega \) is written as

$$\begin{aligned} d\omega =\sum _{i<j}\left( \frac{\partial a_{j}(x)}{\partial x_{i}}-\frac{\partial a_{i}(x)}{\partial x_{j}} \right) dx_{i}\wedge dx_{j}. \end{aligned}$$

From Stokes’ theorem, the following equation holds true:

$$\begin{aligned} \int _{M}d\omega =\int _{\partial M} \omega . \end{aligned}$$
(A-1)

We define smooth consumption surfaces by first defining smooth price surfaces because it is easy to define smooth price surfaces first and then define smooth consumption surfaces using smooth price surfaces in a roundabout way. Given a price pair \((p^{1, 1}, p^{2, 2})\in \mathbb {R}_{++}^{N} \times \mathbb {R}_{++}^{N}\) and the two price bundles \((p^{1, 2}, p^{2, 1})\in \mathbb {R}_{++}^{N} \times \mathbb {R}_{++}^{N}\) satisfying \(p^{1, 2}=V^{p}(p^{2, 2})\cap R(p^{1, 1})\) and \(p^{2, 1}=V^{p}(p^{1, 1})\cap R(p^{2, 2})\), we consider smooth price surfaces with boundary surrounded by the four prices \((p^{1, 1}, p^{1, 2}, p^{2, 1}, p^{2, 2})\). Let \(S^{p}(p^{1, 1}, p^{2, 2})\) be a smooth price surface in an N-dimensional price space with boundary generated by a price pair \((p^{1, 1}, p^{2, 2})\in \mathbb {R}_{++}^{N} \times \mathbb {R}_{++}^{N}\) that satisfies the following three conditions:

  1. (1)

    \(S^{p}(p^{1, 1}, p^{2, 2})\) is surrounded by the four prices \((p^{1, 1}, p^{1, 2}, p^{2, 1}, p^{2, 2})\) and accompanying four boundary curves.

  2. (2)

    The first and second boundary curves are two iso-price-ratio curves from \(p^{1, 1}\) to \(p^{1, 2}\) and \(p^{2, 1}\) to \(p^{2, 2}\):

    $$\begin{aligned} \{ p\in \mathbb {R}_{++}^{N}|p\in R(p^{1, 1}), U(x(p^{2, 2}))\ge U(x(p)) \ge U(x(p^{1, 1})) \} \end{aligned}$$

    and

    $$\begin{aligned} \{ p\in \mathbb {R}_{++}^{N}|p\in R(p^{2, 2}), U(x(p^{2, 2}))\ge U(x(p)) \ge U(x(p^{1, 1})) \}, \end{aligned}$$

    respectively.

  3. (3)

    The third and fourth boundary curves are a pair of two indifference price paths, which are denoted as \(C_{p}^{D}(i; p^{1, 1}, p^{2, 1})\) from \(p^{1, 1}\) to \(p^{2, 1}\) and \(C_{p}^{D}(i; p^{1, 2}, p^{2, 2})\) from \(p^{1, 2}\) to \(p^{2, 2}\), that satisfy

    $$\begin{aligned} C_{p}^{D}(i; p^{1, 2}, p^{2, 2})\in R(q), \;\;q=C_{p}^{D}(i; p^{1, 1}, p^{2, 1}), \forall i. \end{aligned}$$

These four curves are chosen so that the integrations along the four curves are closely related to the compensating and equivalent variations.Footnote 33 Define \(S^{x}(p^{1, 1}, p^{2, 2})=\{x\in \mathbb {R}_{+}^{N}| x=x(p), p\in S^{p}(p^{1, 1}, p^{2, 2}) \}\) as a smooth consumption surface in an N-dimensional consumption space with boundary generated by a price pair \((p^{1, 1}, p^{2, 2})\in \mathbb {R}_{++}^{N} \times \mathbb {R}_{++}^{N}\). Given a smooth consumption surface \(S^{x}(p^{1, 1}, p^{2, 2})\), we consider a smooth consumption curve on the smooth consumption surface \(S^{x}(p^{1, 1}, p^{2, 2})\):

$$\begin{aligned} C_{x}(i; x^{1, 1}, x^{2, 2})\in S^{x}(p^{1, 1}, p^{2, 2}),\;s\le i \le t, \end{aligned}$$

where \(x^{1, 1}=x(p^{1, 1})\) and \(x^{2, 2}=x(p^{2, 2})\).

1.2 A.2 Propositions and lemmas

Assume that the prices change from \(p^{1, 1}\) to \(p^{2, 2}\). Consider a smooth consumption surface \(S^{x}(p^{1, 1}, p^{2, 2})\).Footnote 34 In order to calculate surface integrals, we need to choose two tangent vectors on the smooth consumption surface at each point. We take \(x_{w}(p(x), w)\) as the first tangent vector on the smooth consumption surface \(S^{x}(p^{1, 1}, p^{2, 2})\) at x.Footnote 35\(x_{w}(p, w)\) is clearly one of the tangent vectors on the smooth consumption surface \(S^{x}(p^{1, 1}, p^{2, 2})\), because, under no money illusion, a change in the wealth level is equivalent to the corresponding proportional price change, that is, a price change along an iso-price-ratio curve.

Let \(c(x)=(c_{1}, c_{2}, c_{3}, \ldots , c_{N})\in \mathbb {R}^{N}\) be the second tangent vector that satisfies the following conditions:

  1. (1)

    At each consumption bundle x on \(S^{x} (p^{1, 1}, p^{2, 2})\), there are two directions to which x can move while keeping it still on both \(S^{x} (p^{1, 1}, p^{2, 2})\) and the indifference consumption hypersurface \(V^{x} (x)\). Let \(C_{x}(i; y, z)\) be a consumption curve from \(y \in R(p^{1, 1})\) via x to \(z \in R(p^{2, 2})\) each point of which is located on both \(S^{x}(p^{1, 1}, p^{2, 2})\) and \(V^{x}(x)\). I choose the direction of c(x) such that c(x) and \(\partial C_{x}(i; y, z) / \partial i \) have the same direction, where where \(x=C_{x}(i; y, z)\). Intuitively speaking, the direction of c(x) is chosen such that it points from \(R(p^{1, 1})\) to \(R(p^{2, 2})\) while keeping it on both \(S^{x} (p^{1, 1}, p^{2, 2})\) and \(V^{x} (x)\).

  2. (2)

    \(|c(x)|=1\).

By construction, \(c(x)\cdot p(x)=0\) holds true. In sum, we take \(x_{w}(p(x), w)\) and c(x) as the two tangent vectors at each point on the smooth consumption surface \(S^{x}(p^{1, 1},p^{2, 2})\).

We take the smooth consumption surface \(S^{x}(p^{1, 1}, p^{2, 2})\) as a two dimensional smooth submanifold M with boundary.Footnote 36 The boundary \(\partial M\) can be divided into two consumption curves from \(x^{1, 1}=x(p^{1, 1})\) to \(x^{2, 2}=x(p^{2, 2})\). Let \(C_{x}^{1}(i; x^{1, 1}, x^{2, 2})\) be the consumption curve from \(x^{1, 1}\) via \(x^{2, 1}=x(p^{2, 1}), p^{2, 1}=V^{p}(p^{1, 1})\cap R(p^{2, 2})\) to \(x^{2, 2}\) on the boundary \(\partial M\), and let \(C_{x}^{2}(i; x^{1, 1}, x^{2, 2})\) be the consumption curve from \(x^{1, 1}\) via \(x^{1, 2}=x(p^{1, 2}), p^{1, 2}=V^{p}(p^{2, 2})\cap R(p^{1, 1})\) to \(x^{2, 2}\) on the boundary \(\partial M\): \(C^{j}_{x}(i; x^{1, 1}, x^{2, 2})\in \partial M, \forall i, j\).

Proposition 9

Given a smooth consumption surface \(S^{x}(p^{1, 1}, p^{2, 2})\), the following holds true:

$$\begin{aligned} \varDelta CS_{1} - \varDelta CS_{2}=\int _{S}\tau (x)dS, \end{aligned}$$
(A-2)

where

$$\begin{aligned} \tau (x)= & {} -\frac{1}{\mu (c(x), x_{w}(p, w))} \sum _{i = 1}^{N} \sum _{j = i}^{N}c_{i} (x)\\&\times \left( \frac{\partial p_{i} (x, w)}{\partial x_{j}} - \frac{\partial p_{j} (x, w)}{\partial x_{i}} \right) \frac{\partial x_{j}(p, w)}{\partial w} \\ \varDelta CS_{1}= & {} \int _{C_{x}^{1}}\sum _{i=1}^{N}p_{i}(x)dx_{i},\;\varDelta CS_{2}=\int _{C_{x}^{2}}\sum _{i=1}^{N}p_{i}(x)dx_{i}, \\ \mu (c(x), x_{w}(p, w))= & {} \sqrt{|c(x)|^{2}|x_{w}(p, w)|^{2}-(c(x)\cdot x_{w}(p, w))^{2}}, \end{aligned}$$

and we integrate from \(x^{1, 1}\) to \(x^{2, 2}\).

Proof

When the value of a 1-form \(\omega \) on \(M=S^{x}(p^{1, 1}, p^{2, 2})\) at x is written as \(\omega _{x}=\sum _{i=1}^{N}p_{i}(x)dx_{i}\),Footnote 37 the integration of the 1-form \(\omega \) on the boundary \(\partial M\) is written as:Footnote 38

$$\begin{aligned} \int _{\partial M} \omega =\varDelta CS_{1} - \varDelta CS_{2}, \end{aligned}$$
(A-3)

where we integrate from \(x^{1, 1}\) to \(x^{2, 2}\). \(\varDelta CS_{1}\) and \(\varDelta CS_{2}\) can be considered as the changes in consumer surplus along the respective integration paths. Thus, the right side of Eq. (A-3), \(\varDelta CS_{1} - \varDelta CS_{2}\), can be considered as the difference between the change in consumer surplus along \(C_{x}^{1}\) and that along \(C_{x}^{2}\). From Appendix A.3, the integration of \(d\omega \), denoted by \(\int _{M}d\omega \), is written as

$$\begin{aligned}&\int _{S}\frac{1}{\mu (c(x), x_{w}(p, w))} \nonumber \\&\quad \sum _{i<j}\left( \frac{p_{j}(x, w)}{\partial x_{i}}- \frac{p_{i}(x, w)}{\partial x_{j}} \right) \left( \frac{\partial x_{j}(p, w)}{\partial w}c_{i}(x)- \frac{\partial x_{i}(p, w)}{\partial w}c_{j}(x) \right) dS. \end{aligned}$$
(A-4)

Arranging Eq. (A-4) and using Eq. (A-1), we get

$$\begin{aligned} \int _{M}d\omega = -\int _{S}\frac{1}{\mu (c(x), x_{w}(p, w))} \sum _{i = 1}^{N} \sum _{j = i}^{N} c_{i} (x) \left( \frac{\partial p_{i} (x, w)}{\partial x_{j}} - \frac{\partial p_{j} (x, w)}{\partial x_{i}} \right) \frac{\partial x_{j}(p, w)}{\partial w} d S \end{aligned}$$

Thus, we get Proposition 9. \(\square \)

Let \(\mathscr {C}_{x}(S^{x}(p^{1, 1}, p^{2, 2}))\) be the set of all the smooth consumption curves from \(x^{1, 1}\) to \(x^{2, 2}\) on a smooth consumption surface \(S^{x}(p^{1, 1}, p^{2, 2})\). We can generalize Proposition 9 as follows.

Corollary 1

Consider a submanifold \(M_{1} \subset M=S^{x}(p^{1, 1}, p^{2, 2})\) whose boundary \(\partial M_{1}\) consists of two smooth consumption curves \(C_{x}^{1, i}\in \mathscr {C}(S^{x}(p^{1, 1}, p^{2, 2})), i=1, 2\) that do not intersect each other. The difference between the change in consumer surplus along \(C_{x}^{1, 1}\) and that along \(C_{x}^{1, 2}\) is written as

$$\begin{aligned} \varDelta CS_{1, 1}-\varDelta CS_{1, 2}&=\int _{M_{1}}\tau (x)dS, \end{aligned}$$

where

$$\begin{aligned} \varDelta CS_{1, 1}=\int _{C_{x}^{1, 1}}\sum _{i=1}^{N}p_{i}(x)dx_{i},\quad \varDelta CS_{1, 2}=\int _{C_{x}^{1, 2}}\sum _{i=1}^{N}p_{i}(x)dx_{i}, \end{aligned}$$

and we integrate from \(x^{1, 1}=x(p^{1, 1})\) to \(x^{2, 2}=x(p^{2, 2})\).

Proof

This is proven in the same way as in the proof of Proposition 9. \(\square \)

Let \(C_{x}^{0}(i; x^{1, 1}, x^{2, 2})\) be the observed consumption path. Along the consumption path \(C_{x}^{0}(i; x^{1, 1}, x^{2, 2})\), the change in consumer surplus can be calculated. Consider the smooth consumption surface \(S^{x}(p^{1, 1}, p^{2, 1}), p^{1, 1}=p(x^{1, 1}), p^{2, 2}=p(x^{2, 2})\) that contains all points of the consumption path \(C_{x}^{0}(i; x^{1, 1}, x^{2, 2})\). By using Proposition 4 and Corollary 1, the compensating and equivalent variations are calculated from the observed change in consumer surplus if we know the Marshallian demand function.

The scalar field \(\tau (x)\) can be expressed in different ways.

Lemma 1

The scalar field \(\tau (x)\) can be rewritten as

$$\begin{aligned} \tau (x)&= \frac{1}{\mu (c(x), x_{w}(p, w))}\sum _{i = 1}^{N} \sum _{j = 1}^{N} c_{i} (x) \frac{\partial p_{j} (x, w)}{\partial x_{i}} \frac{\partial x_{j} (p, w)}{\partial w}\\&=-\frac{1}{\mu (c(x), x_{w}(p, w))} \sum _{i = 1}^{N} \sum _{j = 1}^{N} c_{i} (x) \frac{\partial p_{i} (x, w)}{\partial x_{j}} x_{j}. \end{aligned}$$

Proof

From Mas-Colell et al. (1995),

$$\begin{aligned} p_{i} + \sum _{j = 1}^{N} \frac{\partial p_{i} (x, w)}{\partial x_{j}}\frac{\partial x_{j} (p, w)}{\partial w} = 0, \quad p_{i} + \sum _{j = 1}^{N} \frac{\partial p_{j} (x, w)}{\partial x_{i}} x_{j} =0 \end{aligned}$$

hold true.Footnote 39 Observe that

$$\begin{aligned} \sum _{i = 1}^{N} \sum _{j = 1}^{N} c_{i} (x) \frac{\partial p_{i} (x, w)}{\partial x_{j}} \frac{\partial x_{j} (p, w)}{\partial w} = 0, \quad \sum _{i = 1}^{N} \sum _{j = 1}^{N} c_{i} (x) \frac{\partial p_{j} (x, w)}{\partial x_{i}} x_{j} = 0 \end{aligned}$$

hold true. Here, if

$$\begin{aligned} \sum _{i = 1}^{N} \sum _{j = 1}^{N} c_{i} (x) \left( \frac{\partial p_{i} (x, w)}{\partial x_{j}} -\frac{\partial p_{j} (x, w)}{\partial x_{i}} \right) \left( \frac{\partial x_{j} (p, w)}{\partial \ w} - x_{j} \right) =0 \end{aligned}$$
(A-5)

hold true, then Lemma 1 must hold true. Thus, we prove Eq. (A-5) instead. When \(x_{w}(p, w)-x=0\), it is obvious. The other cases are proven as follows: \(x_{w}(p, w)\) can be written as \(x_{w}(p, w)=x+c\), where c is a column vector and orthogonal to p because the inner product of p and both sides of \(x_{w}(p, w)=x+c\) are written as \(p\cdot x_{w}(p, w)=p\cdot (x+c)\Longrightarrow 1=1+0=1\). Thus, \(x_{w}(p, w)-x\) is orthogonal to p unless x and \(x_{w}(p, w)\) are the same. Consider a smooth surface S with boundary on an indifference consumption hypersurface. S can be considered a two-dimensional smooth submanifold M. Assume that the boundary of M, \(\partial M\), consists of a simple closed curve C. When the value of a 1-form \(\omega \) on M at x is written as \(\omega _{x}=\sum _{i=1}^{N}p_{i}(x)dx_{i}\), the integration of the 1-form \(\omega \) on the boundary \(\partial M\) is written as \(\int _{\partial M}\omega =0\) because the smooth surface M is on an indifference consumption hypersurface; thus, the change in consumer surplus is zero. The integration of the exterior derivative of \(\omega \) is written asFootnote 40

$$\begin{aligned} \int _{M}d\omega = \int _{S}A \sum _{i = 1}^{N }\sum _{j = 1}^{N} c_{i}^{1} (x) \left( \frac{\partial p_{i} (x, w)}{\partial x_{j}} -\frac{\partial p_{j} (x, w)}{\partial x_{i}} \right) c_{j}^{2} (x) dS=0, \end{aligned}$$
(A-6)

where A is a scalar, and \(c^{1}(x)\) and \(c^{2}(x)\) are tangent vectors at x on the smooth surface M and are not parallel. By substituting c(x) for \(c^{1}(x)\) and \(x_{w}(p, w)-x\) for \(c^{2}(x)\) and factoring in the fact that all derivatives of the inverse demand function are continuous, Lemma 1 is proven. \(\square \)

Corollary 2

Given a smooth surface \(S=S^{x}(p^{1, 1}, p^{2, 2})\), the following holds true:

$$\begin{aligned} \frac{r(p^{2, 1}, p^{2, 2})}{r(p^{1, 1}, p^{1, 2})}&=\exp \left[ \int _{S}\tau (x)dS \right] . \end{aligned}$$

Proof

From Propositions 1 and 2, the changes in consumer surplus, \(\varDelta CS_{1}\) and \(\varDelta CS_{2}\), are written as \(\varDelta CS_{1}=\log (r(p^{2, 1}, p^{2, 2}))\) and \(\varDelta CS_{2}=\log (r(p^{1, 1}, p^{1, 2}))\). Exponentiating the both sides of Proposition 9, we get Corollary 2. \(\square \)

1.3 A.3 Calculation of Eq. (A-4)

When a basic two form \(dx_{i}\wedge dx_{j}\) accepts as input two vectors \(v^{1}\in \mathbb {R^{N}}\) and \(v^{2}\in \mathbb {R^{N}}\), the output is written as

$$\begin{aligned} dx_{i}\wedge dx_{j}(v^{1}, v^{2})=\det \left[ \begin{array}{cc} dx_{i}(v^{1})&{} dx_{i}(v^{2}) \\ dx_{j}(v^{1})&{} dx_{j}(v^{2}) \end{array} \right] =\det \left[ \begin{array}{cc} v^{1}_{i}&{}v^{2}_{i} \\ v^{1}_{j}&{}v^{2}_{j} \end{array} \right] =v^{1}_{i}v^{2}_{j}-v^{1}_{j}v^{2}_{i}, \end{aligned}$$

where \(v^{1}_{i}\) and \(v^{2}_{i}\) are the i-th elements of \(v^{1}\) and \(v^{2}\) and \(v^{1}_{j}\) and \(v^{2}_{j}\) are the j-th elements of \(v^{1}\) and \(v^{2}\), respectively. Thus, the integration of \(d\omega \) is written as

$$\begin{aligned} \int _{M}d\omega= & {} \int _{S}\frac{1}{\mu (c(x), x_{w}(p, w))}d\omega (c(x), x_{w}(p, w)) \\= & {} \int _{S}\frac{1}{\mu (c(x), x_{w}(p, w))}\sum _{i<j}\left( \frac{\partial p_{j}(x, w)}{\partial x_{i}}-\frac{\partial p_{i}(x, w)}{\partial x_{j}} \right) dx_{i}\wedge dx_{j}(c(x), x_{w}(p, w))\\= & {} \int _{S}\frac{1}{\mu (c(x), x_{w}(p, w))}\sum _{i<j}\left( \frac{\partial p_{j}(x, w)}{\partial x_{i}}-\frac{\partial p_{i}(x, w)}{\partial x_{j}} \right) \det \left[ \begin{array}{cc} c_{i}(x)&{}\frac{\partial x_{i}(p, w)}{\partial w} \\ c_{j}(x)&{}\frac{\partial x_{j}(p, w)}{\partial w} \end{array} \right] dS \\= & {} \int _{S}\frac{1}{\mu (c(x), x_{w}(p, w))}\sum _{i<j}\left( \frac{p_{j}(x, w)}{\partial x_{i}}- \frac{p_{i}(x, w)}{\partial x_{j}} \right) \\&\times \left( \frac{\partial x_{j}(p, w)}{\partial w}c_{i}(x)- \frac{\partial x_{i}(p, w)}{\partial w}c_{j}(x) \right) dS, \end{aligned}$$

where \(\mu (c(x), x_{w}(p, w))\) is the normalization parameter that normalizes the area of a parallelogram formed by c(x) and \(x_{w}(p, w)\) to one, that is, \(\mu (c(x), x_{w}(p, w))\) is the reciprocal of the area of a parallelogram formed by c(x) and \(x_{w}(p, w)\).

1.4 A.4 Proof of Proposition 7

We need to know \(\tau (x)\) at \(x^{1, 1}\) and the area of \(S^{x}(p^{1, 1}, p^{2, 2})\) to prove Eq. (6). First, \(c(x^{1, 1})\) can be written as a linear combination of \(x_{w}^{1, 1}\) and \(\varDelta x\): \(c(x^{1, 1})=a_{1}x_{w}^{1, 1}+a_{2}\varDelta x\), where \(a_{1}\) and \(a_{2}\) are scalars. The inner product of \(c(x^{1, 1})\) and \(p^{1, 1}\) is \(p^{1, 1}\cdot c(x^{1, 1})=a_{1}+a_{2}p^{1, 1}\cdot \varDelta x=0\). Thus, we get \(a_{1}=-a_{2}(p^{1, 1}\cdot \varDelta x)\). Then, part of \(\tau (x^{1, 1})\) is calculated asFootnote 41

$$\begin{aligned}&\sum _{i = i}^{N}\sum _{j = 1}^{N} \frac{\partial x_{i}^{1, 1} }{\partial w}\frac{\partial p_{i}^{1, 1} }{\partial x_{j}} c_{j} = \sum _{i = i}^{N}\sum _{j = 1}^{N} \frac{\partial x_{i}^{1, 1} }{\partial w}\frac{\partial p_{i}^{1, 1} }{\partial x_{j}} \left( a_{1} \frac{\partial x_{j}^{1, 1} }{\partial w} + a_{2} \varDelta x_{j} \right) \\&\quad = \sum _{i = 1}^{N} \frac{\partial x_{i}^{1, 1}}{\partial w} ( - a_{1} p_{i}^{1, 1} + a_{2} \varDelta p_{i} ) = a_{2} (p^{1, 1}\cdot \varDelta x+\varDelta p \cdot x_{w}^{1, 1} ) \\&\quad =a_{2}\varDelta p\cdot (x_{w}^{1, 1}-x^{1, 1} ), \end{aligned}$$

where \(\partial x_{i}^{1, 1}/ \partial w =\partial x_{i} (p^{1, 1}, w) / \partial w \), \(\partial p_{i}^{1, 1} / \partial x_{j} = \partial p_{i} (x^{1, 1}, w) / \partial x_{j}\), and

$$\begin{aligned} \sum _{j = 1}^{N} \frac{\partial p_{i}^{1, 1} }{\partial x_{j}} \frac{\partial x_{j}^{1, 1} }{\partial w} = -p_{i}^{1, 1},\quad \sum _{j = 1}^{N} \frac{\partial p_{i}^{1, 1} }{\partial x_{j}} \varDelta x_{j} = \varDelta p_{i}, \quad \sum _{i = 1} \frac{\partial x_{i}^{1, 1}}{\partial w} p_{i}^{1, 1} = 1. \end{aligned}$$

Next, \(\varDelta x\) can be written as a linear combination of \(x_{w}(p, w)\) and c(x):

$$\begin{aligned} \varDelta x=b_{1}x_{w}^{1, 1}+b_{2}c(x^{1, 1})=(b_{1}-a_{2}b_{2}(p^{1, 1}\cdot \varDelta x))x_{w}^{1, 1}+a_{2}b_{2}\varDelta x, \end{aligned}$$

where \(b_{1}\) and \(b_{2}\) are scalars. Thus, \(a_{2}b_{2}=1\) and \(b_{1}=(p^{1, 1}\cdot \varDelta x)\) must hold true. This means that \(b_{1}=p^{1, 1}\cdot \varDelta x\) and \(b_{2}= 1 / a_{2}\). The area of \(S^{x}(p^{1, 1}, p^{2, 2})\) is written as

$$\begin{aligned} S=b_{1}b_{2}\sqrt{|c(x^{1, 1})|^{2}|x_{w}^{1, 1}|^{2}-(c(x^{1, 1})\cdot x_{w}^{1, 1})^{2}}=b_{1}b_{2}\mu (c(x^{1, 1}), x_{w}^{1, 1}). \end{aligned}$$

Thus, the difference between the change in consumer surplus along \(C_{x}^{1}(i; x^{1, 1}, x^{2, 2})\) and that along \(C_{x}^{2}(i; x^{1, 1}, x^{2, 2})\) is written as

$$\begin{aligned} \int _{S}\tau (x)dS&=b_{1}b_{2}\mu (c(x^{1, 1}), x_{w}^{1, 1})\times \frac{1}{\mu (c(x^{1, 1}), x_{w}^{1, 1})}a_{2}\varDelta p\cdot (x_{w}^{1, 1}-x^{1, 1})\\&=-\ (\varDelta p\cdot x^{1, 1})\{ \varDelta p\cdot (x_{w}^{1, 1}-x^{1, 1}) \}. \end{aligned}$$

When \(\varDelta p\) is small enough, the area surrounded by \(C_{x}^{1}(i; x^{1, 1}, x^{2, 2})\) and \(C_{x}^{2}(i; x^{1, 1}, x^{2, 2})\) can be considered as a parallelogram. Since the straight line \(C_{0}(i; x^{1, 1}, x^{2, 2})\) from \(x^{1, 1}\) to \(x^{2, 2}\) divides the parallelogram into two triangles whose areas are the same, Eqs. (7) and (8) hold true.

Appendix B: Calculation of Eq. (23)

Choose \(p^{i, i}, \forall i\) such that \(\varDelta CS_{0, i}=\varDelta CS_{0, j}\) holds true for all i and j. When M is large enough, \(\log \left( r(p^{i+1, i}, p^{i+1, i+1})\right) \) can be approximated by \(\varDelta CS_{0, i}\). Thus, \(\varDelta CS_{1}\) is approximately written as

$$\begin{aligned} \varDelta CS_{1}\approx \sum _{i=1}^{M-1} \varDelta CS_{0, i} \prod _{j=1}^{M-i-1} (1-\eta ^{i+j, i})= \sum _{i=1}^{M-1} \frac{\varDelta CS_{0}}{M-1} \prod _{j=1}^{M-i-1} (1-\eta ^{i+j, i}). \end{aligned}$$

When \(\eta ^{i, j}, \forall i, j\) are constant and all the same, the above equation is rewritten as

$$\begin{aligned} \varDelta CS_{1}\approx \sum _{i=1}^{M-1}(1-\eta )^{i}\frac{\varDelta CS_{0}}{M-1}, \end{aligned}$$
(A-7)

where \(\eta =\eta ^{i, j}, \forall i, j\). Since \(\varDelta p^{i+1, i} \approx \varDelta p^{j, j+1}, \forall i, j\) holds true, \(\eta \) is approximated by \(( 1+ \alpha )^{M-1}-1\), where \(\alpha =\varDelta p \cdot (x_{w}^{1, 1}-x^{1, 1})\). Thus, when M increases, Equation (A-7) is calculated asFootnote 42

$$\begin{aligned} \varDelta CS_{1}\approx \lim _{z \rightarrow \infty } \sum _{i=1}^{z}\left\{ 2-( 1+ \alpha )^{\frac{1}{z}} \right\} ^{i}\frac{\varDelta CS_{0}}{z}=\frac{\alpha }{\alpha +1}\frac{\varDelta CS_{0}}{\log (1+\alpha )}. \end{aligned}$$

Using the same reasoning, \(\varDelta CS_{2}\) is calculated as \(\alpha (\varDelta CS_{0} / \log (1+\alpha ))\).

Appendix C: Application to price indices

1.1 C.1 Relationship between welfare evaluations and price indices

This subsection shows the relationships between price indices and measures of consumer welfare change. The following relationships have been already indicated or pointed out by Hicks (1941), Hulten (1973), Bruce (1977), and others. However, it is important to formally state the following relationships between price indices and measures of welfare change to make it easier to understand the main results in this paper. Assume that the prices and the consumption bundle change from \((p^{1}, x^{1})\) at time 1 to \((p^{2}, x^{2})\) at time 2, where \(p^{1}\cdot x^{1}=w^{1}\) and \(p^{2}\cdot x^{2}=w^{2}\) hold. When these changes are caused by a policy change, the change in economic welfare can be evaluated by using consumer surplus, compensating variation, or equivalent variation. Also, the price and quantity indices at time 1 and 2 can be calculated.

Let \(P_{i}(p^{1}, p^{2}, x^{1}, x^{2})\) be the true cost of living price index, and let \(Q_{i}(p^{1}, p^{2}, x^{1}, x^{2})\) be the corresponding implicit quantity index. All the price and quantity indices at time 1 are assumed to be normalized to one. The Laspeyres–Konüs true cost of living price index is written as

$$\begin{aligned} P_{L}(p^{1}, p^{2}, x^{1}, x^{2})= \frac{e(p^{2}, v(p^{1}, w^{1}))}{e(p^{1}, v(p^{1}, w^{1}))}, \end{aligned}$$
(A-8)

and the Paasche–Konüs true cost of living price index is written as

$$\begin{aligned} P_{P}(p^{1}, p^{2}, x^{1}, x^{2})= \frac{e(p^{2}, v(p^{2}, w^{2}))}{e(p^{1}, v(p^{2}, w^{2}))}. \end{aligned}$$
(A-9)

The Laspeyres–Konüs true cost of living price index and its corresponding implicit quantity index, denoted by \(Q_{L}(p^{1}, p^{2}, x^{1}, x^{2})\), satisfy the following equation:

$$\begin{aligned} P_{L}(p^{1}, p^{2}, x^{1}, x^{2})Q_{L}(p^{1}, p^{2}, x^{1}, x^{2})=\frac{w^{2}}{w^{1}}. \end{aligned}$$
(A-10)

The Paasche–Konüs true cost of living price index and its corresponding implicit quantity index, denoted by \(Q_{P}(p^{1}, p^{2}, x^{1}, x^{2})\), satisfy the following equation:

$$\begin{aligned} P_{P}(p^{1}, p^{2}, x^{1}, x^{2})Q_{P}(p^{1}, p^{2}, x^{1}, x^{2})=\frac{w^{2}}{w^{1}}. \end{aligned}$$
(A-11)

When income level w is normalized to one, the followings hold:

$$\begin{aligned}&CV=1-e(p^{2}, v(p^{1}, w)),\quad EV=e(p^{1}, v(p^{2}, w))-1 \\&P_{L}(p^{1}, p^{2}, x^{1}, x^{2})= e(p^{2}, v(p^{1}, w)), \quad P_{P}(p^{1}, p^{2}, x^{1}, x^{2})= \frac{1}{e(p^{1}, v(p^{2}, w))}\\&P_{L}(p^{1}, p^{2}, x^{1}, x^{2})Q_{L}(p^{1}, p^{2}, x^{1}, x^{2})= P_{P}(p^{1}, p^{2}, x^{1}, x^{2})Q_{P}(p^{1}, p^{2}, x^{1}, x^{2})=1. \end{aligned}$$

Thus, rearranging the above equations, we get the following equations:

$$\begin{aligned} CV&=1-P_{L}(p^{1}, p^{2}, x^{1}, x^{2})=1-\frac{1}{Q_{L}(p^{1}, p^{2}, x^{1}, x^{2})} \end{aligned}$$
(A-12)
$$\begin{aligned} EV&= \frac{1}{P_{P}(p^{1}, p^{2}, x^{1}, x^{2})} -1=Q_{P}(p^{1}, p^{2}, x^{1}, x^{2})-1. \end{aligned}$$
(A-13)

Next, we consider the relationship between the Divisia price index and the change in consumer surplus. Let \(P_{D}(i)\) be the Divisia price index at time i. Assume that the prices and the consumption bundle continuously change from \((p^{1}, x^{1})\) at time \(t=1\) to \((p^{2}, x^{2})\) at time \(s=2\) along a price path \(C_{p}(i, p^{1}, p^{2}), s\le i \le t\), and that income level w is normalized to one at each point along \(C_{p}(i, p^{1}, p^{2})\). The Divisia price index at time 2 is written asFootnote 43

$$\begin{aligned} P_{D}(2)=\exp \left[ \int _{s}^{t}\frac{\sum _{j=1}^{N} x_{j}(q(i), w )dq_{j}(i)}{\sum _{j=1}^{N} x_{j}(q(i), w )q_{j}(i)} \right] =\exp \left[ \int _{s}^{t}\sum _{j=1}^{N} x_{j}(q(i), w )dq_{j}(i) \right] . \end{aligned}$$
(A-14)

From Eqs. (1) and (A-14), we get

$$\begin{aligned} P_{D}(2)=\exp (-\varDelta CS). \end{aligned}$$
(A-15)

Thus, when income level w is normalized to one at each time along a price path, the compensating variation is convertible with the Laspeyres–Konüs true cost of living price index and its corresponding implicit quantity index by using Eq. (A-12), the equivalent variation is convertible with the Paasche–Konüs true cost of living price index and its corresponding implicit quantity index by using Eq. (A-13), and the change in consumer surplus and the Divisia price index along the price path are convertible with each other by using Eq. (A-15). Therefore, even though we mainly deal with consumer surplus, compensating variation, and equivalent variation in the main text, the results provided in the main text are also applicable to these price and quantity indices.

1.2 C.2 Price indices

Under the assumption of Proposition 8, the Laspeyres–Konüs true cost of living price index and the Paasche–Konüs true cost of living price index are written as

$$\begin{aligned} P_{L}(p^{1, 1}, p^{2, 2}, x^{1, 1}, x^{2, 2})=\frac{1}{\exp \left( \varDelta CS_{1} \right) },\quad P_{P}(p^{1, 1}, p^{2, 2}, x^{1, 1}, x^{2, 2})=\frac{1}{\exp \left( \varDelta CS_{2} \right) }. \end{aligned}$$

In large price change cases in Sect. 2.5, the Laspeyres–Konüs true cost of living price index and the Paasche–Konüs true cost of living price index are written as

$$\begin{aligned}&P_{L}(p^{1, 1}, p^{M, M}, x^{1, 1}, x^{M, M})\approx \frac{1}{\exp \left( \frac{\alpha }{\alpha +1}\frac{\varDelta CS_{0}}{\log (1+\alpha )} \right) } \end{aligned}$$
(A-16)
$$\begin{aligned}&P_{P}(p^{1, 1}, p^{M, M}, x^{1, 1}, x^{M, M})\approx \frac{1}{\exp \left( \alpha \frac{\varDelta CS_{0}}{\log (1+\alpha )} \right) }. \end{aligned}$$
(A-17)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sasaki, T. Welfare evaluations and price indices with path dependency problems. Soc Choice Welf 52, 127–159 (2019). https://doi.org/10.1007/s00355-018-1142-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00355-018-1142-4

Navigation