Skip to main content

Generalized Goodwin’s Theorems on General Coordinates

  • Chapter
  • First Online:
Complexities of Production and Interacting Human Behaviour
  • 597 Accesses

Abstract

This paper is concerned with providing foundations for the use of general coordinates in linear production systems, which Professor R. M. Goodwin originally proposed. The actually observed sectors of the Sraffa–Leontief system have bilateral relations to each other, which leads to complexities in analyzing prices or quantities. In the case that it is impossible to proceed with analysis directly in the original space, we may turn to an imaginary space, which is tractable, as long as the original space corresponds uniquely to the imaginary one adopted. This is a natural way for mathematics to proceed. Goodwin’s use of the method of general coordinates represents a “diffeomorphisrn” in this area of economic theory. The use of general coordinates in the Sraffa–Leontief system seems, at first glance, restrictive, since their use requires a diagonalizable input matrix. A mathematical meaning of diagonalizability (or the rank condition) is examined by means of linear perturbation. The eigenvalues of the linear perturbed system are seen to be distinct almost everywhere. Economic meaning is given to the diagonalizable input matrix as yielding a regular system. It is worth noting that this specification can be justified in essentially the same way as the notion of quasi-smoothness used by Mas-Colell in the neoclassical production set. As a by-product of this study, the perturbed version of the Sraffa–Leontief price system is analysed.

Reprinted from Structural Change and Economic Dynamics 2(1), Aruka, Y., Generalized Goodwin Theorems on General Coordinates, 69–91 (1991), http://www.sciencedirect.com/science/journal/0954349X. With kind permission from Elsevier, Oxford, UK.

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 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
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

Notes

  1. 1.

    See Goodwin (1976, 1983b, c) and Goodwin and Punzo (1987). General coordinates are often called “principal coordinates” by Goodwin. See, for example, Goodwin and Punzo (1987, p. Chap. 2.5: Principal Coordinates).

  2. 2.

    In the following, [a i diag] indicates a diagonal matrix whose diagonal elements are a i .

  3. 3.

    See, for example, Nikaido (1968, p. 103, Theorem 7.2 iv), Seneta (1981, pp 22–23). As for essentially non-negative matrices, see Kato (1982, p. 146. Theorem 7.2).

  4. 4.

    In the original quotation, Goodwin considered each eigenvalue being eligible for a measure of capital intensity. He also confused a necessary condition for an increase of a principal eigenvalue with a sufficient condition for an increase of capital intensity. That is to say, he carelessly wrote: “If μ i , increases, more inputs per unit are required”.

  5. 5.

    [Sraffa] defines a standard commodity, which is the dominant eigenvector of his linear production system. But there are n − 1 other output eigenvectors (and eigenvalues), and it proves illuminating to deploy all of them instead of merely the only purely real and positive one (Goodwin 1986, p. 205).

  6. 6.

    See Strang (1980, p. 235) and also Aruka (1987, Sect. 11). “For any square matrix, there is a unitary matrix S = U such that \( {U^{ - 1}}TU = T \) is upper triangular” (Strang 1980, p. 234). This proposition is called “Schur’s Lemma”. If a unitary matrix satisfies that A Hermitian is equal to A Inverse, namely \( {A^H} = {A^{ - 1}} \), the upper triangularization implies diagonalization. In the world of mutually adjoint (Hermitian conjugate) matrices, diagonalization is always guaranteed (Strang 1980).

  7. 7.

    Blatt suggested an introduction of a slight adjustment to obtain desired result on the invertibility on the matrix. “[If] an irreducible matrix has no inverse, changing a few coefficients by infinitesimal amounts is enough to make it have an inverse; hence the conditions are not seriously restrictive” (Blatt 1983, p. 118).

  8. 8.

    See e.g. Kan and Iri (1982, Chap. 5), Kato (1982. Chap. 2) and also Wilkinson (1965, Chap. 2).

  9. 9.

    “The production set Y is quasi-smooth if there is a closed region of Lehesgue measure zero Z such that the corresponding distance function C 2 is on \( {R^n}\backslash Z \) (Mas-Colell 1985, p. 107, 3.6.1.)”. As for the close similarity between our diagonality assumption and Mas-Colell’s quasi-smoothness, see Aruka (1990).

  10. 10.

    The weighted mean of the μ-group eigenvalues of A(ε) with the multiplicity h is

    $$ (1/h){\rm trace}[A(\varepsilon )P(\varepsilon )] = \mu + (1/h){\rm trace}\{ [A(\varepsilon ) - \mu I]P(\varepsilon )]\}. $$
  11. 11.

    A coefficient system when being faced with perturbation seems inherently unstable. Aruka (1988) discussed the uncertainty principle in the simultaneous adjustment system of price and quantity underlying the Sraffa–Leonticf system.

  12. 12.

    See e.g. Strang (1980, p. 193). It is also noted that if AB = BA and both A and B can be diagonalized, then they can be diagonalized by the same H.

  13. 13.

    Goodwin failed to formulate an analytically transformed system such as (2.14), with which eigenvectors are to be integrated as a general solution. However, each particular solution e mr u is connected with each other through a common rate of profit r. We replace (2.13) with the next one where wages are paid in advance:

    p = (1+r)(Ap+a 0).

    Transforming this system onto general coordinates, it follows

    υ i  = (1+r)(μ i υ i +b i ).

    Here p = Hu, a 0 = Hb. Thus

    μ i +b i (1/υ i ) = 1/(1+r).

    This is the relationship that Goodwin preferred to use (see Goodwin, 1983a, pp. 157–158; Goodwin-Punzo, 1987. p. 92).

  14. 14.

    Similar statements of this theorem (i) were originally discussed in Schefold (1976) and Miyao (1977). However, they did not state the equivalence of this theorem to Theorem 1.

  15. 15.

    See Wilkinson (1965, pp. 88–89) and Strang (1980, pp. 284–285). As for a further analysis of perturbation, see e.g. Koyarik and Sherif (1985).

  16. 16.

    It is verified that our regular production system is structurally stable under perturbation. It is also shown that a system around the neighborhood of a given regular system is regular against the same primary factor. See Theorem 3 in Aruka (1990).

  17. 17.

    p(0) is a solution of (2.12) when r is set to zero:

    p(0) = [IA] a 0.

    This is nothing but a vector of labor values.

  18. 18.

    A study of this issue had been done in the context of a Nash equilibrium elsewhere. See Aruka (1989).

References

  • Aruka Y (1987) Generalized Goodwin’s theorem in the linear multisectoral model. Research Papers No. 9. The Institute of Business Research, Chuo University (Presented at the Far Eastern Meeting of the Econometric Society, Tokyo, Japan)

    Google Scholar 

  • Aruka Y (1988) General coordinates and price dynamics in the multisectoral model. Research Papers No. 11. The Institute of Business Research, Chuo University (Presented at the Australasian Meeting of the Econometric Society, Canberra, Australia)

    Google Scholar 

  • Aruka Y (1989) Perturbed version of the Leontief price system and the Nash equilibrium. Research Papers No. 13. The Institute of Business Research, Chuo University

    Google Scholar 

  • Aruka Y (1990) Perturbation theorems on the linear production model and some properties of eigenprices. Working Paper in Economics and Econometrics, Australian National University, No. 203

    Google Scholar 

  • Benavie A (1972) Mathematical techniques for economic analysis. Prcntice-Hall, Englewood Cliffs, NJ

    Google Scholar 

  • Blatt JM (1983) Dynamic economic systems. M.E. Sharpe, New York

    Google Scholar 

  • Dietzenbacher E (1988) Perturbations of matrices: a theorem on the Perron vector and its applications to input-output models. J Econ 48:389–412

    Google Scholar 

  • Gantmacher FR (1959) The theory of matrices, vol I. Chelsea, New York

    Google Scholar 

  • Goodwin RM (1953) Static and dynamic linear general equilibrium models. In: Netherlands Economic Institute (ed) Input–output relations. Netherlands Economic Institute, The Netherlands (also reprinted in Goodwin 1983a)

    Google Scholar 

  • Goodwin RM (1976) Use of normalized general coordinates in linear value and distribution theory. In: Polenske K and Skolka J (eds) Advances in input–output analysis (also reprinted in Goodwin 1983a)

    Google Scholar 

  • Goodwin RM (1983a) Essays in linear economic structures. Macmillan, London

    Google Scholar 

  • Goodwin RM (1983b) Capital theory in orthogonalised general coordinates. In: Essays in linear economic structures. Macmillan, London

    Google Scholar 

  • Goodwin RM (1983c) Disaggregating models of fluctuating growth. In: Goodwin RM, Krueger M, Vercelli A (eds) Nonlinear models of fluctuating growth. Springer, Berlin

    Google Scholar 

  • Goodwin RM (1986) Swinging along the Turnpike with von Neumann and Sraffa. Camb J Econ 10:203–210

    Google Scholar 

  • Goodwin RM, Punzo LF (1987) The dynamics of a capitalist society. Polity, Basil Blackwell, Cambridge

    Google Scholar 

  • Kan T, Iri M (1982) Jordan Hyojunkei (Jordan Normal Forms). University of Tokyo Press, Tokyo (in Japanese)

    Google Scholar 

  • Kato T (1982) A short introduction to perturbation theory for linear operators. Springer, NewYork

    Book  Google Scholar 

  • Klein E (1973) Mathematical methods in theoretical economics. Academic, New York

    Google Scholar 

  • Koyarik ZV, Sherif N (1985) Perturbation of invariant subspaces. Linear Algebra Appl 64:93–113

    Article  Google Scholar 

  • Mas-Colell A (1985) The theory of general economic equilibrium: a differential approach. Cambridge University Press, New York

    Google Scholar 

  • McKenzie L (1960) Matrices with dominant diagonals and economic theory. In: Arrow KJ, Karlln S, Suppes P (eds) Mathematical methods in the social sciences 1959. Stanford University Press, Stanford

    Google Scholar 

  • Miyao T (1977) Generalization of Sraffa’s standard commodity and its complete characterization. Int Econ Rev 18:151–162

    Article  Google Scholar 

  • Nikaido H (1968) Convex structures and economic theory. Academic, New York

    Google Scholar 

  • Satake I (1974) Senkei Daisugaku (Linear algebra). Shokabo, Tokyo (in Japanese)

    Google Scholar 

  • Schefold B (1976) Relative prices as a function of the rate of profit. Zeitschrift fuer Nationaloekonomie 36:21–48

    Article  Google Scholar 

  • Schefold B (1978) On counting equations. Zeitschrift fuer Nationaloekonomie 38:253–285

    Article  Google Scholar 

  • Seneta E (1981) Non-negative matrices and Markov chains, 2nd edn. Springer, New York

    Google Scholar 

  • Sraffa P (1960) Production of commodities by means of commodities. Cambridge University Press, Cambridge

    Google Scholar 

  • Strang G (1980) Linear algebra and its applications, 2nd edn. Academic, Orlando

    Google Scholar 

  • Wilkinson JH (1965) The algebraic eigenvalue problem. Clarendon, Oxford

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yuji Aruka .

Editor information

Editors and Affiliations

Appendix

Appendix

Proof of Proposition 3

Let μ i be eigenvalues of a square matrix A of order n. Then we have a triangularization:

$$ {H^{ - 1}}AH = [{\mu_i}{{\ }_{diag}}]. $$

We also define such a diagonal matrix G that its diagonal elements are g ii =τ i(i =1, …, n). Here we choose any τ fulfilling

$$ 0 \,<\, n\tau \,<\, {\min_{{\mu_i} \ne {\mu_j}}}\left| {{\mu_i} - {\mu_j}} \right|. $$

If μ 1 = μ 2 =  … = μ n, we may choose τ arbitrarily. Furthermore, define F = HGH −1. Thus it follows that

$$ {H^{ - 1}}AH + \varepsilon G = {H^{ - 1}}[A + \varepsilon F]H = [{\mu_i} + \varepsilon {g_{ii}}{{\ }_{diag}}] $$

for a sufficiently small ε ≠ 0. The diagonal elements by definition are all distinct. Hence, any matrix can be made to have all distinct eigenvalues through fine tuning □

Proof of Theorem 1

Since the components of R(g, g′), a i , b i are determined by ε and f kl . R(g, g′) can be expressed a polynomial of ε and f kl . Arranging this polynomial with respect to ε for some   mn 2 −1,

$$ {\upsilon_0}({\,f_{kl}}){\varepsilon^m} + {\upsilon_1}({\,f_{kl}}){\varepsilon^{m - 1}} + \cdots + {\upsilon_{m - 1}}({\,f_{kl}})\varepsilon + {\upsilon_m}({\,f_{kl}}) = 0. $$

Since ε is a sufficiently small non-zero number, R(g, g′) = 0 holds only if

$$ {\upsilon_0}({\,f_{kl}}) = 0,{\upsilon_1}({\,f_{kl}}) = 0,...,{\upsilon_{m - 1}}({\,f_{kl}}) = 0,{\upsilon_m}({\,f_{kl}}) = 0. $$

These are the condition that A + εF has multiple roots for a sufficiently small non-zero number ε. These are not identically fulfilled with any perturbation direction F, which, as a whole, constitutes a projective space of dimension n 2 −1. Only a perturbation direction F specified by the above condition gives multiple roots. Thus such a perturbation direction determines an algebraic manifold V of a smaller dimension than n 2 −1 in a whole projective space. Excepting these singularities, eigenvalues of A + εF are all distinct for any perturbation direction F. A graph of V is of measure zero. Hence we have all distinct eigenvalues almost everywhere.□

Proof of Corollary 1 to Theorem 1

\( {\upsilon_m}({\,f_{kl}}) \) is really det|A|. If A has no repeated roots, det|A| does not vanish. Hence \( {\upsilon_m}({\,f_{kl}}) \) does not vanish. A + εF then will not have repeated roots for a sufficiently small ε(0). □

Proof of Theorem 3

In matrix the differential equation system is of the form:

$$ \frac{{dp}}{{dr}} = Ap + (1 + r)A\frac{{dp}}{{dr}}. $$
(2.25)

This is reformulated into the following form:

$$\begin{array}{lll} \frac{{dp}}{{dr}} = \{ {[I - (1 + r)A]^{ - 1}}A\} p \\ \quad = \{ {[I - (1 + r)A]^{ - 1}}{({A^{ - 1}})^{ - 1}}\} p. \\ \quad = {\{ {A^{ - 1}}[I - (1 + r)A]\}^{ - 1}}p \cr \quad = {[{A^{ - 1}} - (1 + r)I]^{ - 1}}p. \\\end{array} $$

By the assumptions, \( [I - (1 + r)A] \) is non-negatively invertible because of the Frobenius Theorem. Then it is expanded in the following series:

$$ {[I - (1 + r)A]^{{ - }1}} = I + (1 + r)A + {(1 + r)^2}{A^2} +... $$
(2.26)

Employing the expansion, it is verified that \( {[I - (1 + r)A]^{ - 1}}A \) and A are commutable:

$$ {[I - (1 + r)A]^{ - 1}}A \cdot A = A \cdot {[I - (1 + r)A]^{ - 1}}A\ \hbox{for}\ r \le R. $$
(2.27)

Therefore, by virtue of Lemma matrix \( {[I - (1 + r)A]^{ - 1}}A \) has the same eigenvectors with a commutable matrix A. We denote \( {[I - (1 + r)A]^{ - 1}}A \) by B. Let matrix H be constructed by the eigenvectors u of matrix A. The general solution of this differential equation system can then be of the formFootnote 16:

$$ p(r) = {e^{Br}}p(0). $$

If A is non-singular, it follows that

$$ {H^{ - {1}}}AH = [{\mu_i}{_{\ diag}}], $$

from which it is easily transformed:

$$ {A^{ - 1}} = {H^{ - 1}}{[{\mu_i}{_{\,diag}}]^{ - 1}}H = {H^{ - 1}}\left[ {\frac{1}{{{\mu_i}}}{_{diag}}} \right]H. $$
(2.28)

By making use of (2.15) and (2.28), the similarity transformation of \( {[I - (1 + r)A]^{ - 1}}A \) is derived in the following manner:

$$ \begin{array}{lll} {[I - (1 + r)A]^{ - 1}}A = {[{A^{ - 1}} - (1 + r)I]^{ - 1}} \\ \qquad\qquad\qquad\qquad = {\left[ {H\left[ {\frac{1}{{{\mu_i}}}{_{diag}}} \right]{H^{ - 1}} - (1 + r)I} \right]^{ - 1}} \cr \qquad\qquad\qquad\qquad = H\left[ {\left[ {\frac{1}{{{\mu_i}}}{_{diag}}} \right] - (1 + r)I} \right]^{ - 1}{H^{ - 1}}. \\\end{array} $$

Accordingly,

$$ {H^{ - 1}}{[I - (1 + r)A]^{ - 1}}AH = {\left[ {\frac{1}{{{\mu_i}}} - (1 + r){_{\,diag}}} \right]^{ - 1}} = M. $$
(2.29)

The general solution then is represented as

$$ p(r) = {e^{Br}}p(0) = H{e^{Mr}}{H^{ - 1}}p(0) = \sum\limits_{i = 1}^M {{e^{{m_i}r}}} {c_i}{u_i}, $$

where \( M = [{m_i}{{ }_{\,diag}}] = {\left[ {\frac{1}{{{\mu_i}}} - (1 + r){{ }_{diag}}} \right]^{ - 1}}, \) and c i , is a constant determined by the initial condition c = H –1 p(0).

Applying p =e mr u to (2.25) as a particular solution, it follows that

$$ mp = {[{A^{ - 1}} - (1 + r)I]^{ - 1}}p. $$

This is arranged as follows:

$$ Ap = \frac{m}{{1 + m(1 + r)}}p. $$

By replacing \( \frac{m}{{1 + m(1 + r)}} \) with μ,

$$ Ap = \mu p. $$

Thus it has been proved that the eigenvalues m of the differential system correspond to the eigenvalues μ of the Sraffa–Leontief price system by the relationship of \( \frac{\mu }{{1 - \mu (1 + r)}} \). □

Proof of Corollary 2

In the assumptions of Theorem 3. the principal eigenvalue is only one real value. Choose r closely to the maximum rate of profit R defined by the principal eigenvalue \( \mu (A) = {{1 + R}}. \) Then, in the principal eigensector, the corresponding component of M in (2.29) will be:

$$ {\left\{ {\frac{1}{{\mu (A)}} - (1 + r)} \right\}^{ - 1}} = \frac{{\mu (A)}}{{1 - \mu (A)(1 + r)}} \,> 0, $$

since \( \frac{1}{{1 + r}} \,>\, \mu (A) \,>\, 0. \) The particular solution of the principal eigensector will definitely be divergent. In a productive system, all the eigenvalues are smaller than 1 in terms of absolute values:

$$ 1 > |{\mu_i}|. $$

Consequently, the general solution will be unstable as the rate of profit becomes the maximum rate of profit R. □

Proof of Corollary 3

In our price movement equation system the initial value is the value when r is set to 0:

$$ p(0) = {[I - A]^{ - 1}}{a^0}, $$

which can be expanded in the following manner, given the assumptions of Theorem 3:

$$ p(0) = {a^0} + A{a^0} + {A^2}{a^0} +... \,.$$

Taking account of our requirement Aa 0 = μa 0, this will furthermore be rearranged into

$$ p(0) = (1 + \mu + {\mu^2} +...){a^0} = \frac{1}{{1 - \mu }}{a^0}. $$

Then it follows that

$$ p(r) = \frac{1}{{1 - \mu }}H{e^{Mr}}{H^{ - 1}}{a^0}. $$

H is constructed by the columns of eigenvectors u of A. Substitute a 0 into u 1. The initial value will be expressed as follows:

$$ c = {H^{ - 1}}{a^0} = [1,0,...,0]\prime, $$

because of H −1 H = I. Hence other coefficients other than the first component of c are 0.

Therefore

$$ p(r) = \sum\limits {{e^{{m_i}r}}} {c_i}{u_i} = {e^{{m_1}}}{u_1} = {e^{{m_1}}}{a^0}. $$

This completes the proof. □

Proof of Theorem 4

Let m be the ratio satisfying the condition

$$ m = \frac{{{a_i}p}}{{a_i^0}} = \frac{{{a_i}Ap}}{{{a_i}{a^0}}} = \frac{{{a_i}{A^2}p}}{{{a_i}A{a^0}}} =...{\hbox{ for }}r \in (0,R]. $$

This ratio implies “the balancing proportion” for the prices against the changes of the rate of profit. The activity vector s preserving this critical proportion has to satisfy the following condition:

$$ m = \frac{{s{A^n}p}}{{{a_i}{A^{n - 1}}{a^0}}}\ \hbox{for}\ r \in (0,R]. $$
(2.30)

Substituting (2.12) into (2.30), it follows that, taking account of (2.27),

$$ s\{ {[I - (1 + r)A]^{ - 1}}A - mI\} {A^{n - 1}}{a^0} = 0. $$

Let z be the vector \( s\{ {[I - (1 + r)A]^{ - 1}}A - mI\} \) and L the matrix \( [{a^0},A{a^0},{A^2}{a^0},...,{A^{n - 1}}{a^0}] \):

$$ zL = 0. $$
(2.31)

Hence if L is of dimension n, the null space of L is of dimension nn =0. Thus (2.31) has only the solution z = 0. Conversely, if z = 0, the dimension of L is n. L of dimension n implies that

$$ {a^0},A{a^0},{A^2}{a^0},...,{A^{n - 1}}{a^0} $$

are linearly independent. z = 0 means that s is a left-hand eigenvector of matrix \( {[I - (1 + r)A]^{ - 1}}A \):

$$ s{[I - (1 + r)A]^{ - 1}}A = ms, $$

i.e.

$$ s[{A^{ - 1}} - (1 + r)A] = ms. $$

This completes the desired results:

\( sA = s \frac {m} {1 +\,m(1+r)}\) for a left-hand eigenvector s,

\( Au = \frac {m} {1 +\,m(1+r)}\) for a right-hand eigenvector u.

The second result is just the same as Theorem 3(ii). Thus the necessary and sufficient condition of this theorem is equivalent to normal modes of the Sraffa–Leontief differential equation system (2.14).

Finally, it remains to examine the assumptions sa 0 ≠ 0, and a 0 ≠ (1 + R)Aa 0. It is easily verified that if sa 0 ≠ 0, a 0 ≠ (1 + R)Aa 0, then \( {a^0},A{a^0},{A^2}{a^0},...,{A^{n - 1}}{a^0} \) are linearly independent. Suppose, on the contrary, they are linearly dependent. In the case n = 2, appropriating the non-zero numbers 1 and −(1 + R), it follows that a 0−(1 + R)Aa 0=0. a 0 happens to be a left-hand eigenvector of A. Then, by orthogonality of the left- and right-hand eigenvectors in an eigenvector space, sa 0 = 0. Conversely, if sa 0 = 0, a 0 is an eigenvector of A. While z ≠ 0. In the case n = 2, then, in view of Jordan normal forms, it can be expressed:

$$ {s^1} = m{s^1} + {s^2}. $$

Here s i is a generalized eigenvector. Post-multiplying the left-hand vector a 0 on both sides, it follows that

$$ {s^1}{a^0} = m{s^1}{a^0} + {s^2}{a^0}. $$

This implies s 2 a 0 is 0 if s 1 a 0 is 0. By induction on a cyclic subspace, it is proved that if a 0, \( {a^0},A{a^0},{A^2}{a^0},...,{A^{n - 1}}{a^0} \)are linearly independent, sa 0 ≠ 0 (rsp. a 0 ≠ (1+R)Aa 0). □

Proof of Proposition 4

By one of the Frobenius theorems, the principal eigenvalues \( \mu (A), \widetilde{\mu }(\tilde{A}) \) fulfilling \( Ap = \mu (A)p, \widetilde{A}\widetilde{p} = \widetilde{\mu }(\widetilde{A})\widetilde{p} \) are positive and simple. Suppose, on the contrary, that for \( \mu (A) > \widetilde{\mu }(\widetilde{A}) \),

$$ \frac{{{{\widetilde{p}}_k}}}{{{p_k}}} > \frac{{\mu (A)}}{{\widetilde{\mu }(\widetilde{A})}}ma{x_{j \in \Gamma }}\frac{{{{\widetilde{p}}_j}}}{{{p_j}}}. $$

Then it follows that

$$ \frac{{\mu (A)}}{{\widetilde{\mu }(\widetilde{A})}} > ma{x_{j \in \Gamma }}\frac{{{{\widetilde{p}}_j}/{p_j}}}{{{{\widetilde{p}}_k}/{p_k}}} > 1. $$

Hence \( \mu (A) < \widetilde{\mu }(\widetilde{A}) \). This is a contradiction. Similar to the above, we can prove:

$$ \frac{{{{\widetilde{p}}_k}}}{{{p_k}}} \ge \frac{{\mu (A)}}{{\widetilde{\mu }(\widetilde{A})}}mi{n_{j \in \Gamma }}\frac{{{{\widetilde{p}}_j}}}{{{p_j}}}\ \hbox{for}\ \mu (A) < \widetilde{\mu }(\widetilde{A}). $$

Proof of Theorem 5

  1. (i)

    If it is the case that

    $$ \mu (A) + \varepsilon a \ > \frac{1}{{1 + r}} > \mu (A), $$

    a price system will be made unreasonable for a given rate of profit r. In other words, the reasonableness of a price system after perturbation will depend on not only a matrix F but also on the magnitude of a given rate of profit. Thus Proposition 5 supports statement (i).Footnote 17

  1. (ii)

    The perturbed matrix A + εF is assumed to be non-singular. Define H(ε) to be a transformation matrix being constructed by the eigenvectors u(ε) of A + εF. All the arguments in Theorem 3 will be valid, if A is replaced with A + εF as well as μ with μ + εa. Hence

    $$ \begin{array}{lll} p(r) = H(\varepsilon ){e^{M(\varepsilon )r}}H{(\varepsilon )^{ - 1}}p(0) \\ \qquad = \sum\limits_{i = 1}^M {{e^{{m_i}(\varepsilon )r}}} {c_i}(\varepsilon ){u_i}, \\\end{array} $$

    where

    $$ M(\varepsilon ) = [m{(\varepsilon )_i}{_{\,diag}}] = {\left[ {{{\left\{ {\frac{1}{{\mu {{(\varepsilon )}_i}}} - (1 + r)} \right\}}_{diag}}} \right]^{ - 1}}, $$

    c(ε) is a constant determined by the initial condition c(ε)= H(ε)−1 p(0). □

Rights and permissions

Reprints and permissions

Copyright information

© 2011 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Aruka, Y. (2011). Generalized Goodwin’s Theorems on General Coordinates. In: Aruka, Y. (eds) Complexities of Production and Interacting Human Behaviour. Physica, Heidelberg. https://doi.org/10.1007/978-3-7908-2618-0_2

Download citation

Publish with us

Policies and ethics