Skip to main content

A Theory for Estimating Consumer’s Preference from Demand

  • Chapter
Advances in Mathematical Economics Volume 19

Part of the book series: Advances in Mathematical Economics ((MATHECON,volume 19))

Abstract

This study shows that if the estimate error of a demand function satisfying the weak axiom of revealed preference is sufficiently small with respect to local C 1 topology, then the estimate error of the corresponding preference relation (which is possibly nontransitive, but uniquely determined from demand function, and transitive under the strong axiom) is also sufficiently small. Furthermore, we show a similar relation for the estimate error of the inverse demand function with respect to the local uniform topology. These results hold when the consumption space is the positive orthant, but are not valid in the nonnegative orthant.

JEL Classification: D11.

Mathematics Subject Classification (2010): 91B16, 62P20.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 54.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. Afriat S (1967) The construction of a utility function from demand data. Int Econ Rev 8:67–77

    Article  Google Scholar 

  2. Antonelli GB (1886) Sulla Teoria Matematica Dell’ Economia Politica. Tipografia del Folchetto, Pisa

    Google Scholar 

  3. Debreu G (1972) Smooth preferences. Econometrica 40:603–615

    Article  MathSciNet  Google Scholar 

  4. Hartman P (1997) Ordinary differential equations. Society for Industrial and Applied Mathematics, Birkhaeuser Verlag AG

    Google Scholar 

  5. Hildenbrand W (1974) Core and equilibria of a large economy. Princeton University Press, New Jersey

    Google Scholar 

  6. Hosoya Y (2013) Measuring utility from demand. J Math Econ 49:82–96

    Article  MathSciNet  Google Scholar 

  7. Hurwicz L, Uzawa H (1971) On the integrability of demand functions. In: Chipman JS, Hurwicz L, Richter MK, Sonnenschein HF (eds) Preferences, utility and demand. Harcourt Brace Jovanovich, New York, pp 114–148

    Google Scholar 

  8. Kannai Y (1970) Continuity properties of the core of a market. Econometrica 38:791–815

    Article  Google Scholar 

  9. Kim T, Richter MK (1986) Nontransitive-nontotal consumer theory. J Econ Theory 34:324–363

    Article  MathSciNet  Google Scholar 

  10. Pareto V (1906) Manuale di economia politica con una introduzione alla scienza sociale. Societa Editorice Libraria, Milano

    Google Scholar 

  11. Pontryagin LS (1962) Ordinary differential equations. Addison-Wesley (translated from Russian). Reading, Massachusetts

    Google Scholar 

  12. Quah JK-H (2006) Weak axiomatic demand theory. Econ Theory 29:677–699

    Article  MathSciNet  Google Scholar 

  13. Richter MK (1966) Revealed preference theory. Econometrica 34:635–645

    Article  Google Scholar 

  14. Samuelson PA (1950) The problem of integrability in utility theory. Economica 17:355–385

    Article  MathSciNet  Google Scholar 

  15. Smale S, Hirsch MW (1974) Differential equations, dynamical systems, and linear algebra. Academic, New York

    Google Scholar 

Download references

Acknowledgements

We are grateful to Toru Maruyama for his helpful comments and suggestions. We would also like to express gratitude to the anonymous referee for their helpful advice.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yuhki Hosoya .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Proof of Theorem 1

We at first introduce a lemma.

Lemma 1.

Let \((\succsim _{k})_{k}\) be a sequence on \(\mathcal{P}\) and \(\succsim \in \mathcal{P}\). Suppose

$$ \displaystyle\begin{array}{rcl} & x \succsim _{k}v \Leftrightarrow u_{k}(x,v) \geq 1& {}\\ & x \succsim v \Leftrightarrow u(x,v) \geq 1 & {}\\ \end{array} $$

for some real-valued function uk,u on \(\Omega ^{2}\). Assume u,uk are continuous and0Note that under these conditions, \(\succsim,\succsim _{k}\) must satisfy completeness and continuity.

$$ \displaystyle\begin{array}{rcl} & u_{k}(x,v) > 1 \Leftrightarrow u_{k}(v,x) < 1,u(x,v) > 1 \Leftrightarrow u(v,x) < 1,& {}\\ & u_{k}(x,av) = \frac{1} {a}u_{k}(x,v),u(x,av) = \frac{1} {a}u(x,v), & {}\\ & u_{k}(x,v) \geq 1\text{ and }u(x,v) \geq 1\text{ if }x \geq v. & {}\\ \end{array} $$

Then, \(\text{Lim}_{k} \succsim _{k} =\succsim \) with respect to the closed convergence topology if \((u_{k})_{k}\) converges to u uniformly on any compact set.

Proof.

Suppose \((u_{k})_{k}\) converges to u uniformly on any compact set, and choose any \((x,v) \in \text{Limsup}_{k} \succsim _{k}\). Then, for any neighborhood U of (x, v) and any \(k_{0} \in \mathbb{N}\), there exists \(k \geq k_{0}\) such that \(\succsim ^{k} \cap U\neq \varnothing \). Therefore, there exists an increasing sequence (k(m)) such that \(u_{k(m)}(x_{m},v_{m}) \geq 1\) and \(\|(x_{m},v_{m}) - (x,v)\| \leq \frac{1} {m}\). By the assumption of uniform convergence, we have \(\vert u_{k(m)}(x_{m},v_{m}) - u(x_{m},v_{m})\vert \rightarrow 0\). Because u is continuous, \(u(x_{m},v_{m}) \rightarrow u(x,v)\), and thus \(u_{k(m)}(x_{m},v_{m}) \rightarrow u(x,v)\). As \(u_{k(m)}(x_{m},v_{m}) \geq 1\), we have u(x, v) ≥ 1, and thus \(\text{Limsup}_{k} \succsim _{k} \subset \succsim \).

Next, under the same assumption, choose any \((x,v) \in \succsim \). Then, we have u(x, v) ≥ 1. This implies \(u(x,(1 - \frac{1} {m})v) > 1\) for any m. Now, take any neighborhood U of (x, v). Then, there exists m > 0 such that \((x,(1 - \frac{1} {m})v) \in U\). Because \(u_{k}(x,(1 - \frac{1} {m})v) \rightarrow u(x,(1 - \frac{1} {m})v) > 1\), there exists \(k_{0} \in \mathbb{N}\) such that \((x,(1 - \frac{1} {m})v) \in \succsim _{k}\) for any k ≥ k 0. Hence, we have \(\succsim \subset \text{Liminf}_{k} \succsim _{k}\). Therefore, we have \(\text{Lim}_{k} \succsim _{k} =\succsim \). ■ 

Hence, it suffices to show that \((u^{g_{k}})_{ k}\) converges to u g uniformly on any compact set.

The next lemma was proved in Hosoya [6]. Thus, we omit the proof of this result.

Lemma 2.

Let \(g: \Omega \rightarrow \mathbb{R}_{++}^{n}\) be C 1 -class, and y(t;x,v) be the nonextendable solution of the following problem:

$$\displaystyle\begin{array}{rcl} \dot{y}& =& (g(y) \cdot x)v - (g(y) \cdot v)x, {}\\ y(0)& =& x, {}\\ \end{array}$$

where \(x,v \in \Omega \) . Define \(w(x,v) = (v \cdot x)v - (v \cdot v)x\) . Then, for any \(z \in \text{span}\{x,v\}\) , z ⋅ w(x,v) = 0 if and only if z is proportional to v. Let ]a,b[ be the domain of \(y(\cdot;x,v)\) . Then, \(t(x,v) =\min \{ t \geq 0\vert y(t;x,v) \cdot w(x,v) \geq 0\}\) is well-defined. Moreover, there exists a continuous function \(y_{1}(x,v),y_{2}(x,v)\) such that \(y_{1}(x,v) = y_{2}(x,v) = x\) if x is proportional to v, and \(y(t(x,v);x,v) \in [y_{1}(x,v),y_{2}(x,v)]\) for any \((x,v) \in \Omega ^{2}\) . Let

$$\displaystyle{A(x,v) =\| x\|\|v -\frac{v \cdot x} {\|x\|^{2}} x\|.}$$

Then, w(x,v) is orthogonal to v, w(x,v) ≠ 0 if and only if x is not proportional to v, and if w(x,v) ≠ 0,

$$\displaystyle{\dot{y} \cdot w(x,v) = A(x,v)^{2}(g(y) \cdot v) > 0.}$$

Hence, y(t;x,v) ⋅ w(x,v) is increasing in t, and thus t(x,v) is a unique t such that y(t;x,v) ⋅ w(x,v) = 0.

Lemma 3.

Suppose \((g_{k})_{k}\) converges to g uniformly on any compact set. Then, \((u^{g_{k}})_{ k}\) converges to u g uniformly on any compact set.

Proof.

Let y(t; x, v) be the nonextendable solution of the following equation:

$$\displaystyle\begin{array}{rcl} \dot{y}& =& (g(y) \cdot x)v - (g(y) \cdot v)x, {}\\ y(0)& =& x, {}\\ \end{array}$$

and y k (t; x, v) be the nonextendable solution of the following equation:

$$\displaystyle\begin{array}{rcl} \dot{y}& =& (g_{k}(y) \cdot x)v - (g_{k}(y) \cdot v)x, {}\\ y(0)& =& x. {}\\ \end{array}$$

Let y(⋅ ; x, v) be defined on ]a(x, v), b(x, v)[. We separate this proof into four steps.

Step 1::

fix any \((x,v) \in \Omega ^{2}\) and c ∈ ]0, b(x, v)[. Then, there exists a compact neighborhood V of (x, v) such that y(⋅ ; y, z) is defined on [0, c] for any (y, z) ∈ V. The set \(\{y(t;y,z)\vert t \in [0,c],(y,z) \in V \}\) is a compact set of \(\Omega \), and thus there exists a compact set \(C \subset \Omega \) such that the above set is included in the interior of C. We shall prove that, for sufficiently large k, \(y_{k}(t;y,z)\) is defined for any \((t,y,z) \in [0,c] \times V\), \(y_{k}(t;y,z) \in C\), and

$$\displaystyle{ \|y_{k}(t;y,z) - y(t;y,z)\| \leq \frac{\|g_{k} - g\|} {M} (e^{2M\|y\|\|z\|t} - 1), }$$
(1)

where M > 0 is a Lipschitz constant of g on C and \(\|g_{k} - g\| =\sup _{w\in C}\|g_{k}(w) - g(w)\|\).

First,

$$\displaystyle\begin{array}{rcl} y_{k}(t;y,z) - y(t;y,z)& =& (y_{k}(t;y,z) - y) - (y(t;y,z) - y) {}\\ & =& \int _{0}^{t}[\dot{y}_{ k}(s;y,z) -\dot{ y}(s;y,z)]ds. {}\\ \end{array}$$

Hence, if t ∈ [0, c] is in the domain of \(y_{k}(\cdot;y,z)\) and \(y_{k}(s;y,z) \in C\) for any s ∈ [0, t], then we have

$$\displaystyle\begin{array}{rcl} & & \|y_{k}(t;y,z) - y(t;y,z)\| \\ & & \qquad \leq \int _{0}^{t}\|g_{ k}(y_{k}(s;y,z)) - g(y(s;y,z))\| \cdot 2\|y\|\|z\|ds \\ & & \qquad \leq \int _{0}^{t}\|g(y_{ k}(s;y,z)) - g(y(s;y,z))\| \cdot 2\|y\|\|z\|ds \\ & & \qquad +\int _{ 0}^{t}\|g_{ k}(y_{k}(s;y,z)) - g(y_{k}(s;y,z))\| \cdot 2\|y\|\|z\|ds \\ & & \qquad \leq 2\|y\|\|z\|\int _{0}^{t}[\|g(y_{ k}(s;y,z)) - g(y(s;y,z))\| +\| g_{k} - g\|]ds \\ & & \qquad \leq 2\|y\|\|z\|\int _{0}^{t}(M\|y_{ k}(s;y,z) - y(s;y,z)\| +\| g_{k} - g\|)ds, {}\end{array}$$
(2)

which implies equation (1).0See Sect. 2.3. Hence, if k is sufficiently large (and thus, \(\|g_{k} - g\|\) is sufficiently small), then \(\|y_{k}(t;y,z) - y(t;y,z)\| \leq \delta\), where δ > 0 is sufficiently small so that the following inequality holds:

$$\displaystyle{\inf _{w\notin C,t\in [0,c],(y,z)\in V }\|w - y(t;y,z)\| \geq 2\delta.}$$

Now, define \(t^{{\ast}} =\sup \{ s \in [0,c]\vert \forall \tau \in [0,s],y_{k}(\tau;y,z) \in C\}\). Then y k (t; y, z) ∈ C for any \(t \in [0,t^{{\ast}}[\). If t is not included in the domain of \(y_{k}(\cdot;y,z)\), then there exists \(\hat{t} < t^{{\ast}}\) such that \(y_{k}(t;y,z)\notin C\) for any \(t \in [\hat{t},t^{{\ast}}[\), a contradiction. Hence, t is in the domain of \(y_{k}(\cdot;y,z)\) and \(y_{k}(t^{{\ast}};y,z) \in C\).

It suffices to show that t  = c. Suppose t  < c. Then, the above arguments implies \(\|y_{k}(t^{{\ast}};y,z) - y(t^{{\ast}};y,z)\| \leq \delta\), and thus \(y_{k}(t^{{\ast}};y,z)\) is included in the interior of C. Hence, y k (⋅ ; y, z) is defined on \([t^{{\ast}},t^{{\ast}}+\varepsilon ]\) and \(y_{k}([0,t^{{\ast}}+\varepsilon ];y,z) \subset C\) for sufficiently small \(\varepsilon > 0\), a contradiction. This completes the proof of our claim.

Step 2::

define w(x, v) as in Lemma 2, and t(x, v) (resp. t k (x, v)) as the minimum constant t ≥ 0 such that y(t; x, v) (resp. y k (t; x, v)) is proportional to v. By Lemma 2, if x is not proportional to v, then t = t(x, v) if and only if y(t; x, v) ⋅ w(x, v) = 0 and t = t k (x, v) if and only if \(y_{k}(t;x,v) \cdot w(x,v) = 0\). Also, again by Lemma 2, t > t(x, v) if and only if y(t; x, v) ⋅ w(x, v) > 0 and t > t k (x, v) if and only if \(y_{k}(t;x,v) \cdot w(x,v) > 0\). Note that w(x, v) = 0 and \(t(x,v) = t_{k}(x,v) = 0\) if and only if x is proportional to v.

Step 3::

choose any compact subset V of \(\Omega ^{2}\). Fix any \(\varepsilon > 0\). Let \(W =\{ (x,v) \in V \vert \|y_{1}(x,v) - y_{2}(x,v)\| <\| v\|\varepsilon \}\), where \(y_{1},y_{2}\) are as in Lemma 2. Because \(y_{1},y_{2}\) is continuous, W is open in V, and thus \(V ^{{\prime}} = V \setminus W\) is compact. By Lemma 2, if (x, v) ∈ W, then both \(y_{k}(t_{k}(x,v);x,v)\) and y(t(x, v); x, v) are included in \([y_{1}(x,v),y_{2}(x,v)]\), and thus,

$$\displaystyle{\|y_{k}(t_{k}(x,v);x,v) - y(t(x,v);x,v)\| \leq \| v\|\varepsilon.}$$

Note that

$$\displaystyle{ u^{g}(x,v) = \frac{1} {\|v\|^{2}}y(t(x,v);x,v) \cdot v,\ u^{g_{k} }(x,v) = \frac{1} {\|v\|^{2}}y_{k}(t_{k}(x,v);x,v) \cdot v. }$$
(3)

Hence, if (x, v) ∈ W,

$$\displaystyle\begin{array}{rcl} \vert u^{g}(x,v) - u^{g_{k} }(x,v)\vert & =& \frac{1} {\|v\|^{2}}\vert (y_{k}(t_{k}(x,v);x,v) - y(t(x,v);x,v)) \cdot v\vert {}\\ &\leq & \frac{\|y_{k}(t_{k}(x,v);x,v) - y(t(x,v);x,v)\|} {\|v\|} {}\\ & \leq &\varepsilon. {}\\ \end{array}$$

Meanwhile, suppose \((x,v) \in V ^{{\prime}}\). If x is proportional to v, then \(y_{1}(x,v) = y_{2}(x,v) = x\), and thus (x, v) ∈ W, a contradiction. Hence, x is not proportional to v. Because t(x, v) is the unique solution of the equation y(t; x, v) ⋅ w(x, v) = 0, by the implicit function theorem, the function t(x, v) is continuous on V . Choose any \((x,v) \in V ^{{\prime}}\) and any \(c(x,v) \in ]t(x,v),b(x,v)[\), and let V (x, v) be a compact neighborhood of (x, v) in V such that \(y(\cdot;y,z)\) is defined on [0, c(x, v)] and t(y, z) < c(x, v) for any \((y,z) \in V _{(x,v)}\). By step 1, we can take a compact set \(C_{(x,v)} \subset \mathbb{R}_{++}^{n}\) such that

  1. (i)

    y(t; y, z) ∈ C (x, v) for any \((t,y,z) \in [0,c(x,v)] \times V _{(x,v)}\),

  2. (ii)

    There exists k 0 such that for any k ≥ k 0 and \((y,z) \in V _{(x,v)}\), y k (⋅ ; y, z) is also defined on [0, c(x, v)] and \(y_{k}(t;y,z) \in C_{(x,v)}\) for any t ∈ [0, c(x, v)], and

  3. (iii)

    For any k ≥ k 0, Eq. (1) holds for any t ∈ [0, c(x, v)].

Because V is compact, there is a finite collection \(\{(x_{1},v_{1}),\ldots,(x_{m},v_{m})\}\) of V such that \(\cup _{i}V _{(x_{i},v_{i})} = V ^{{\prime}}\). Hence, it suffices to show that for any \(i \in \{ 1,\ldots,m\}\), there exists k i such that for any k ≥ k i , \(\sup _{(x,v)\in V _{(x_{ i},v_{i})}}\vert u^{g_{k}}(x,v) - u^{g}(x,v)\vert \leq \varepsilon\). For notational convention, we abbreviate \(V _{(x_{i},v_{i})}\) as V i , \(C_{(x_{i},v_{i})}\) as C i , and \(c(x_{i},v_{i})\) as c i .

Step 4::

by definition of V i , we have t(x, v) is in the domain of \(y_{k}\) if k is sufficiently large. Hence,

$$\displaystyle\begin{array}{rcl} \vert u^{g_{k} }(x,v) - u^{g}(x,v)\vert & =& \frac{1} {\|v\|^{2}}\vert (y_{k}(t_{k}(x,v);x,v) - y(t(x,v);x,v)) \cdot v\vert {}\\ &\leq & \frac{1} {\|v\|^{2}}\vert (y_{k}(t_{k}(x,v);x,v) - y_{k}(t(x,v);x,v)) \cdot v\vert {}\\ & & + \frac{1} {\|v\|^{2}}\vert (y_{k}(t(x,v);x,v) - y(t(x,v);x,v)) \cdot v\vert. {}\\ & & {}\\ \end{array}$$

By Eq. (1),

$$\displaystyle\begin{array}{rcl} & & \frac{1} {\|v\|^{2}}\vert y_{k}(t(x,v);x,v) - y(t(x,v);x,v) \cdot v\vert {}\\ &\leq & \frac{1} {\|v\|}\|y_{k}(t(x,v);x,v) - y(t(x,v);x,v)\| {}\\ & \leq & \frac{\|g_{k} - g\|} {M\|v\|} (e^{2M\|x\|\|v\|c_{i} } - 1), {}\\ \end{array}$$

where the right-hand side is less than \(\frac{\varepsilon }{2}\) uniformly on V i if k is sufficiently large. Hence, it suffices to show that, for any sufficiently large k,

$$\displaystyle{\max _{(x,v)\in V _{i}} \frac{1} {\|v\|^{2}}\vert (y_{k}(t_{k}(x,v);x,v) - y_{k}(t(x,v);x,v)) \cdot v\vert \leq \frac{\varepsilon } {2}.}$$

By the Cauchy-Schwarz inequality, it suffices to show that, for any sufficiently large k,

$$\displaystyle{\max _{(x,v)\in V _{i}}\|y_{k}(t_{k}(x,v);x,v) - y_{k}(t(x,v);x,v)\| \leq \frac{\|v\|\varepsilon } {2}.}$$

Recall that \(y(t;x,v) \cdot w(x,v) > 0\) (resp. \(y_{k}(t;x,v) \cdot w(x,v) > 0\)) if and only if t > t(x, v). (resp. \(t > t_{k}(x,v)\).) As c i  > t(x, v), we have there exists a positive constant C > 0 such that \(y(c_{i};x,v) \cdot w(x,v) \geq C\) for any (x, v) ∈ V i . Hence, for sufficiently large k, \(y_{k}(c_{i};x,v) \cdot w(x,v) > 0\) for any (x, v) ∈ V i . Therefore, \(c_{i} > t_{k}(x,v)\), and thus t k (x, v) is in the domain of y(⋅ ; x, v).

Now, define

$$\displaystyle\begin{array}{rcl} & M_{1} =\inf _{k\in \mathbb{N},(x,v)\in V _{i},y\in C_{i}}\{((g_{k}(y) \cdot x)v - (g_{k}(y) \cdot v)x) \cdot w(x,v)\},& {}\\ & M_{2} =\min _{(x,v)\in V _{i},y\in C_{i}}\{((g(y) \cdot x)v - (g(y) \cdot v)x) \cdot w(x,v)\}. & {}\\ \end{array}$$

By Lemma 2, when x is not proportional to v,

$$\displaystyle\begin{array}{rcl} & ((g_{k}(y) \cdot x)v - (g_{k}(y) \cdot v)x) \cdot w(x,v) = A(x,v)^{2}(g_{k}(y) \cdot v) > 0,& {}\\ & ((g(y) \cdot x)v - (g(y) \cdot v)x) \cdot w(x,v) = A(x,v)^{2}(g(y) \cdot v) > 0. & {}\\ \end{array}$$

Because \(g_{k} \rightarrow g\) uniformly on C i , we can show that \(M_{1},M_{2} > 0\). As both y(t(x, v); x, v) and \(y_{k}(t_{k}(x,v);x,v)\) are proportional to v, we have \(y(t(x,v);x,v) \cdot w(x,v) = y_{k}(t_{k}(x,v);x,v) \cdot w(x,v) = 0\). If \(t(x,v) \geq t_{k}(x,v)\), then

$$\displaystyle\begin{array}{rcl} y_{k}(t(x,v);x,v) \cdot w(x,v)& =& \int _{t_{k}(x,v)}^{t(x,v)}\dot{y}_{ k}(t;x,v) \cdot w(x,v)dt {}\\ & \geq &\int _{t_{k}(x,v)}^{t(x,v)}M_{ 1}dt {}\\ & =& M_{1}(t(x,v) - t_{k}(x,v)). {}\\ \end{array}$$

Hence, we have

$$\displaystyle{t(x,v) \leq t_{k}(x,v) + \frac{1} {M_{1}}y_{k}(t(x,v);x,v) \cdot w(x,v).}$$

Similarly, if \(t(x,v) \leq t_{k}(x,v)\), then

$$\displaystyle\begin{array}{rcl} y(t_{k}(x,v);x,v) \cdot w(x,v)& =& \int _{t(x,v)}^{t_{k}(x,v)}\dot{y}(t;x,v) \cdot w(x,v)dt {}\\ & \geq &\int _{t(x,v)}^{t_{k}(x,v)}M_{ 2}dt {}\\ & =& M_{2}(t_{k}(x,v) - t(x,v)). {}\\ \end{array}$$

Hence, we have

$$\displaystyle{t(x,v) \geq t_{k}(x,v) - \frac{1} {M_{2}}y(t_{k}(x,v);x,v) \cdot w(x,v).}$$

Now, y(t(x, v); x, v) is proportional to v, and thus \(y(t(x,v);x,v) \cdot w(y,z) = 0\). Hence, by Eq. (1), for any c > 0,

$$\displaystyle\begin{array}{rcl} & & \max _{(x,v)\in V _{i}}\vert y_{k}(t(x,v);x,v) \cdot w(x,v)\vert {}\\ &&\qquad =\max _{(x,v)\in V _{i}}\vert (y_{k}(t(x,v);x,v) - y(t(x,v);x,v)) \cdot w(x,v)\vert {}\\ &&\qquad \leq \max _{(x,v)\in V _{i}}\|y_{k}(t(x,v);x,v) - y(t(x,v);x,v)\|\|w(x,v)\| {}\\ & & \qquad \leq c, {}\\ \end{array}$$

for sufficiently large k. Similarly,

$$\displaystyle\begin{array}{rcl} & & \max _{(x,v)\in V _{i}}\vert y(t_{k}(x,v);x,v) \cdot w(x,v)\vert {}\\ &&\qquad =\max _{(x,v)\in V _{i}}\vert (y_{k}(t_{k}(x,v);x,v) - y(t_{k}(x,v);x,v)) \cdot w(x,v)\vert {}\\ &&\qquad \leq \max _{(x,v)\in V _{i}}\|y_{k}(t_{k}(x,v);x,v) - y(t_{k}(x,v);x,v)\|\|w(x,v)\| {}\\ & & \qquad \leq c, {}\\ \end{array}$$

for sufficiently large k. For such k, we have \(\vert t(x,v) - t_{k}(x,v)\vert \leq \frac{c} {M_{3}}\), where \(M_{3} =\min \{ M_{1},M_{2}\}\). Then, if c > 0 is sufficiently small,

$$\displaystyle\begin{array}{rcl} & & \|y_{k}(t_{k}(x,v);x,v) - y_{k}(t(x,v);x,v)\| {}\\ & & \qquad = \left \|\int _{t(x,v)}^{t_{k}(x,v)}\dot{y}_{ k}(t;x,v)dt\right \| {}\\ & & \qquad \leq \left \vert \int _{t(x,v)}^{t_{k}(x,v)}[\max _{ y\in C_{i}}\|(g(y) \cdot x)v - (g(y) \cdot v)x\|]dt\right \vert {}\\ &&\qquad \leq \max _{y\in C_{i}}\|(g(y) \cdot x)v - (g(y) \cdot v)x\| \frac{c} {M_{3}} {}\\ & & \qquad \leq \frac{\|v\|\varepsilon } {2}, {}\\ \end{array}$$

which completes the proof. ■ 

Lemmas 1 and 3 ensure that the claim of Theorem 1 holds. This completes the proof of Theorem 1. ■ 

Appendix 2: Proof of Theorem 2

Define \(h(q) = f(q^{1},\ldots,q^{n-1},1,q^{n})\) and \(h_{k}(q) = f_{k}(q^{1},\ldots,q^{n-1},1,q^{n})\). Then all \(h,h_{k}: \mathbb{R}_{++}^{n} \rightarrow \mathbb{R}_{++}^{n}\) are invertible0That is, all h, h k are bijective. Injectivity follows from the uniqueness of the inverse demand function. Surjectivity arises from the surjectivity and homogeneity of f and f k . and, for any \(q \in \mathbb{R}_{++}^{n}\), there exists a neighborhood U of q such that

$$\displaystyle\begin{array}{rcl} \lim _{k\rightarrow \infty }\sup _{r\in U}\|h_{k}(r) - h(r)\|& =& 0,{}\end{array}$$
(4)
$$\displaystyle\begin{array}{rcl} \lim _{k\rightarrow \infty }\sup _{r\in U}\|Dh_{k}(r) - Dh(r)\|& =& 0.{}\end{array}$$
(5)

Moreover, by the rank condition, we have all Dh(q), Dh k (q) are regular.0See the mathematical appendix of Samuelson [14], or the proof of Proposition 1 in Hosoya [6]. Clearly,

$$\displaystyle{\lim _{k\rightarrow \infty }\|Dh_{k}(q)^{-1} - Dh(q)^{-1}\| = 0}$$

for any \(q \in \mathbb{R}_{++}^{n}\).

Next, choose any \(x \in \Omega \) and corresponding \(q \in \mathbb{R}_{++}^{n}\) such that h(q) = x. Then, there exists \(k_{0} \in \mathbb{N}\) such that some neighborhood of q is included in the domain of h k for any k ≥ k 0. We now introduce a lemma.

Lemma 4.

For any sufficiently small δ > 0, there exists M > 0 such that \(\|h(r) - h(q)\| \leq \delta\) or \(\|h_{k}(r) - h_{k}(q)\| \leq \delta\) for some k ≥ k 0 implies \(\|r - q\| \leq M\delta\) . Moreover, M can be chosen independent to δ.

Proof.

Let

$$\displaystyle\begin{array}{rcl} & \xi (r) = Dh(q)(r - q) - (h(r) - h(q)), & {}\\ & \xi _{k}(r) = Dh_{k}(q)(r - q) - (h_{k}(r) - h_{k}(q)).& {}\\ \end{array}$$

Then, by the assumption of convergence of h k , \(M = 2\sup \{\|Dh(q)^{-1}\|,\) \(\|Dh_{k_{0}}(q)^{-1}\|,\ldots \}< +\infty \). Since \(D\xi (r) = Dh(q) - Dh(r)\) and \(D\xi _{k}(r) = Dh_{k}(q) - Dh_{k}(r)\), for any sufficiently small δ > 0, \(\|r - q\| \leq M\delta\) implies that r is included in the domain of h k for any k ≥ k 0 and0See Eq. (5).

$$\displaystyle{\sup \{\|Dh(q)^{-1}\|\|D\xi (r)\|,\|Dh_{ k_{0}}(q)^{-1}\|\|D\xi _{ k_{0}}(r)\|,\ldots \}\leq \frac{1} {2}.}$$

Now, let \(S =\{ s\vert \|s\| \leq M\delta \}\) and choose any \(r \in \frac{1} {M}S\). Define,

$$\displaystyle{\phi _{r}(s) = Dh(q)^{-1}(r +\xi (s + q)),\phi _{ r}^{k}(s) = Dh_{ k}(q)^{-1}(r +\xi _{ k}(s + q)).}$$

Then, if s ∈ S,

$$\displaystyle\begin{array}{rcl} \|\phi _{r}(s)\|& \leq & \|Dh(q)^{-1}\|\|r\| +\| Dh(q)^{-1}\|\|\xi (s + q)\| {}\\ & \leq & \frac{M} {2} \delta +\| Dh(q)^{-1}\|\|D\xi (ts + q)\|\|s\| {}\\ & \leq & \frac{M} {2} \delta + \frac{1} {2}M\delta \leq M\delta, {}\\ \end{array}$$

where t ∈ [0, 1].0Use the mean value theorem for k(t) = ξ(ts + q). Hence, ϕ r (S) ⊂ S. Moreover,

$$\displaystyle\begin{array}{rcl} \|\phi _{r}(s_{1}) -\phi _{r}(s_{2})\|& \leq & \|Dh(q)^{-1}\|\|\xi (s_{ 1} + q) -\xi (s_{2} + q)\| {}\\ & \leq & \|Dh(q)^{-1}\|\|D\xi (s_{ 3} + q)\|\|s_{1} - s_{2}\| \leq \frac{1} {2}\|s_{1} - s_{2}\|, {}\\ \end{array}$$

where \(s_{3} \in [s_{1},s_{2}]\).0Again, use the mean value theorem for \(k(t) =\xi ((1 - t)s_{1} + ts_{2} + q)\). Hence, ϕ r is a contraction. By contraction mapping fixed point theorem, there uniquely exists s ∈ S such that ϕ r (s) = s, that is,

$$\displaystyle{s = Dh(q)^{-1}r + s - Dh(q)^{-1}(h(s + q) - h(q)),}$$

and thus h(s + q) = h(q) + r. Because h is an injection, we have \(\|h(s + q) - h(q)\| \leq \delta\) implies \(\|s\| \leq M\delta\). Similarly, we have \(\|h_{k}(s + q) - h_{k}(q)\| \leq \delta\) implies \(\|s\| \leq M\delta\). This completes the proof of Lemma 4. ■ 

Note that y = h(r) if and only if \(r = (g^{1}(y),\ldots,g^{n-1}(y),g(y) \cdot y)\) and y = h k (r) if and only if \(r = (g_{k}^{1}(y),\ldots,g_{k}^{n-1}(y),g_{k}(y) \cdot y)\). For any \(\varepsilon > 0\) there exists δ > 0 such that \(\|h(r) - h(q)\| \leq \delta\) or \(\|h_{k}(r) - h_{k}(q)\| \leq \delta\) for some k ≥ k 0 implies \(\|r - q\| \leq M\delta\) and \(M\delta \leq \varepsilon\). Let \(U =\{ y \in \Omega \vert \|x - y\| \leq \delta \}\). Then for any y ∈ U,

$$\displaystyle{\|g_{k}(y) - g_{k}(x)\| \leq \| h_{k}^{-1}(y) - h_{ k}^{-1}(x)\| \leq \varepsilon.}$$

Hence we conclude that \((g_{k})_{k}\) is equicontinuous at x. As x is an arbitrary point of \(\Omega \), we conclude that \((g_{k})_{k}\) is equicontinuous.

Next, we will show that \(g_{k}(x) \rightarrow g(x)\) pointwise. Let \(h(q) = h_{k}(q_{k}) = x\). It suffices to show that \(q_{k} \rightarrow q\). By assumption, there exists a neighborhood V of q and \(k_{0} \in \mathbb{N}\) such that V is included in the domain of h k for any \(k \geq k_{0}\). Without loss of generality, we can assume that V is compact. By assumption, \((h_{k})_{k}\) converges to h uniformly on V and thus, for any δ > 0, there exists \(k_{1} \in \mathbb{N}\) such that \(\|h_{k}(q) - h(q)\| \leq \delta\) for any k ≥ k 1. However, as \(h(q) = x = h_{k}(q_{k})\), we have \(\|h_{k}(q) - h_{k}(q_{k})\| \leq \delta\). Hence, by Lemma 4, if δ > 0 is sufficiently small, then \(\|q - q_{k}\| \leq M\delta\). Thus, \(q_{k} \rightarrow q\).

Therefore, \((g_{k})_{k}\) is equicontinuous and \(g_{k}(x) \rightarrow g(x)\) for any \(x \in \Omega \). Fix any \(x \in \Omega \) and let h(q) = x. Choose any sufficiently small δ > 0 such that \(\|h(r) - h(q)\| \leq \delta\) or \(\|h_{k}(r) - h_{k}(q)\| \leq \delta\) for some k ≥ k 0 implies \(\|r - q\| \leq M\delta\). Then, if \(\|y - x\| \leq \delta\),

$$\displaystyle\begin{array}{rcl} \|g_{k}(y)\|& \leq & \|h_{k}^{-1}(y)\| + 1 {}\\ & \leq & \|h_{k}^{-1}(y) - h_{ k}^{-1}(x)\| +\| h_{ k}^{-1}(x)\| + 1 {}\\ & \leq & M\delta +\sup _{\ell\geq k_{0}}\|h_{\ell}^{-1}(x)\| + 1 < +\infty, {}\\ \end{array}$$

and thus \((g_{k})_{k}\) is uniformly bounded on \(B =\{ y \in \Omega \vert \|y - x\| \leq \delta \}\). If δ > 0 is sufficiently small, then B is compact. Hence, by Ascoli-Arzela theorem, any subsequence of \((g_{k})_{k}\) has a convergent subsubsequence.

Suppose that \((g_{k})_{k}\) does not converge to g uniformly on B. Then, there exists \(\varepsilon > 0\) and a subsequence \((g_{m(k)})_{k}\) such that

$$\displaystyle{\sup _{y\in B}\|g_{m(k)}(y) - g(y)\| \geq \varepsilon.}$$

By the above result, there exists a uniformly convergent subsubsequence \((g_{\ell(k)})_{k}\). Because g k converges to g pointwise, we conclude that

$$\displaystyle{\sup _{y\in B}\|g_{\ell(k)}(y) - g(y)\| <\varepsilon }$$

for some k, a contradiction. This completes the proof. ■ 

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer Japan

About this chapter

Cite this chapter

Hosoya, Y. (2015). A Theory for Estimating Consumer’s Preference from Demand. In: Kusuoka, S., Maruyama, T. (eds) Advances in Mathematical Economics Volume 19. Advances in Mathematical Economics, vol 19. Springer, Tokyo. https://doi.org/10.1007/978-4-431-55489-9_2

Download citation

Publish with us

Policies and ethics