Skip to main content

Bródy’s Stability and Disturbances

  • Chapter
  • First Online:
Modern Classical Economics and Reality

Part of the book series: Evolutionary Economics and Social Complexity Science ((EESCS,volume 2))

  • 782 Accesses

Abstract

Bródy’s conjecture regarding the instability of economies is submitted to an empirical test using input-output flow tables of varying size for the US economy, for the benchmark years 1997 and 2002, as well as for the period 1998–2011. The results obtained lend support to the view of increasing instability of the US economy over the period considered. Furthermore, our analysis shows that only a few vertically integrated industries are enough to shape the behaviour of the entire economy in the case of a disturbance. These results may, on the one hand, provide empirical evidence on the speed of convergence of Marxian iterative procedures ‘transforming’ labour values into production prices; on the other hand, they may usefully be contrasted with those derived in a parallel literature on aggregate fluctuations from microeconomic ‘idiosyncratic’ shocks.

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.

    This chapter draws on Mariolis and Tsoulfidis (2014).

  2. 2.

    Also consider Samuelson (1970) and Hua (1984).

  3. 3.

    See footnote 30 in Chap. 2.

  4. 4.

    For an application of that method to the classical competitive process, which includes simultaneous adjustments in prices and outputs (‘cross-dual dynamics ’), see Egidi (1975). For more recent and general formulations of the cross-dual dynamics, see Steedman (1984), Duménil and Lévy (1989), Halpern and Molnár (1997), Bródy (2000a) and Flaschel (2010, Part 3) and the references therein.

  5. 5.

    It has also been repeatedly conjectured in the literature that, for n tending to infinity, the moduli of the non-dominant eigenvalues (as well as the normalized non-dominant singular values) of random stochastic matrices are uniformly distributed in the interval [0, n −0.5].

  6. 6.

    The SIOTs are provided via the Bureau of Economic Analysis (BEA) website (http://www.bea.gov/iTable/iTable.cfm?ReqID=5&step=1).

  7. 7.

    The two Appendices, at the end of this chapter, briefly extend the analysis to consider the Sraffian multiplier (Appendix 1) and the price effect s of currency devaluation (Appendix 2).

  8. 8.

    Also consider the empirical evidence on \( \left|{\lambda}_{\mathbf{A}2}\right|{\lambda}_{\mathbf{A}1}^{-1} \), from 22 European Union countries, and for the year 2005, where n = 16, 30, 59, provided by Gurgul and Wójtowicz (2015).

  9. 9.

    Also see Molnár and Simonovits (1998), who examine deterministic matrices, and Białas and Gurgul (1998), whose focus is on column stochastic matrices.

  10. 10.

    What follows draws on Mariolis (2008a) and Mariolis and Soklis (2014).

  11. 11.

    For the Keynesian multiplier, see, e.g. Blanchard et al. (2010, Chap. 3); Gnos and Rochon (2008) offer Kaleckian and post-Keynesian explorations of this multiplier. For the multipliers of the traditional input-output analysis , see, e.g. Miller and Blair (2009, Chap. 6) and ten Raa (2005, Chap. 3). Finally, for Marxian versions of the aforesaid multipliers, see, e.g. Lange (1970, Chaps. 2 and 3), Hartwig (2004), Trigg and Philp (2008) and Tsaliki and Tsoulfidis (2015, Chap. 2).

  12. 12.

    For the available input-output data as well as the construction of the relevant variables, see Mariolis and Soklis (2014, Appendix I). Furthermore, Mariolis and Soklis (2014) provide detailed results on the output, import \( \left(\mathbf{p}\widehat{\mathbf{m}}\mathbf{B}{\left[\mathbf{B}-\mathbf{A}\right]}^{-1}\boldsymbol{\Pi} {\mathbf{e}}_i^{\mathrm{T}}\right) \) and employment (eV B Πe T i ) multipliers, and discuss some of their policy implications for the recession-ridden Greek economy, especially for the post-2010 years which are characterized by serious fiscal and external imbalances along with negative net national savings (with the exception of the year 2001, they were negative in each year of the period 2000–2013), negative net investment and exceptionally high unemployment rates (also see Mariolis 2011; Tsoulfidis and Tsaliki 2014).

  13. 13.

    This kind of models is open to serious criticism (see Steedman 2000); the appropriate theoretical framework for dealing with the issue is described in Sects. 2.2.3 and 2.2.4. The above model has been formulated and applied by Mariolis et al. (1997), and the findings were consistent with empirical evidence on the rate of imported cost -inflation in the first year after the last drachma devaluation (by 14 % versus ECU) in March 1998 (the estimated values were in the range of 1.16–1.75 %, while the ‘actual’ one was not considerably greater than 1.2 %; see Bank of Greece 1999, Chap. 4).

  14. 14.

    For the year 1988, the P-F eigenvalues of these matrices are 0.381, 0.939 and 0.821, respectively (Mariolis et al. 1997).

  15. 15.

    A detailed study of nine large post-1990 devaluations (i.e. in excess of 38 % vs. US dollar) shows that the rate of inflation, measured by the consumer price index, is very low relative to the exchange rate devaluation (Burstein et al. 2002). For a combination of the price model (6.14) with Thirlwall’s (1979, 2011) extended model of balance of payments constrained growth , and its application to the Greek economy, for the years 2011–2012, see Mariolis (2014): the findings of that (hypothetical) exercise lend support to the view that a rather large nominal devaluation, i.e. in excess of 57 %–60 %, is a necessary condition for the recovery of the economy.

References

  • Acemoglu, D., Ozdaglar, A., & Tahbaz-Salehi, A. (2010). Cascades in networks and aggregate volatility. National Bureau of Economic Research, Working Paper No. 16516. http://www.nber.org/papers/w16516. Accessed 20 Apr 2015.

  • Acemoglu, D., Carvalho, V. M., Ozdaglar, A., & Tahbaz-Salehi, A. (2012). The network origins of aggregate fluctuations. Econometrica, 80(5), 1977–2016.

    Article  Google Scholar 

  • Aruka, Y. (2015). Evolutionary foundations of economic science. How can scientists study evolving economic doctrines from the last centuries? Tokyo: Springer.

    Google Scholar 

  • Bank of Greece. (1999). Annual report for the year 1998 (in Greek). Athens: Bank of Greece.

    Google Scholar 

  • Berman, A., & Plemmons, R. J. (1994). Nonnegative matrices in the mathematical sciences. Philadelphia: Society for Industrial and Applied Mathematics.

    Book  Google Scholar 

  • Białas, S., & Gurgul, H. (1998). On hypothesis about the second eigenvalue of the Leontief matrix. Economic Systems Research, 10(3), 285–289.

    Article  Google Scholar 

  • Bidard, C., & Schatteman, T. (2001). The spectrum of random matrices. Economic Systems Research, 13(3), 289–298.

    Article  Google Scholar 

  • Blanchard, O., Amighini, A., & Giavazzi, F. (2010). Macroeconomics: A European perspective. London: Prentice Hall.

    Google Scholar 

  • Bródy, A. (1997). The second eigenvalue of the Leontief matrix. Economic Systems Research, 9(3), 253–258.

    Article  Google Scholar 

  • Bródy, A. (2000a). A wave matrix. Structural Change and Economic Dynamics, 11(1–2), 157–166.

    Article  Google Scholar 

  • Bródy, A. (2000b). The monetary multiplier. Economic Systems Research, 12(2), 215–218.

    Article  Google Scholar 

  • Burstein, A., Eichenbaum, M., & Rebelo, S. (2002). Why are rates of inflation so low after large devaluations? National Bureau of Economic Research, Working Paper No. 8748. http://www.nber.org/papers/w8748.pdf. Accessed 12 Dec 2014.

  • Carvalho, V. M. (2014). From micro to macro via production networks. Journal of Economics Perspectives, 28(4), 23–48.

    Article  Google Scholar 

  • Chipman, J. S. (1950). The multi-sector multiplier. Econometrica, 18(4), 355–374.

    Article  Google Scholar 

  • Duménil, G., & Lévy, D. (1989). The competition process in a fixed capital environment: A classical view. The Manchester School, 57(1), 34–57.

    Article  Google Scholar 

  • Dupor, B. (1999). Aggregation and irrelevance in multi-sector models. Journal of Monetary Economics, 43(2), 391–409.

    Article  Google Scholar 

  • Egidi, M. (1975). Stability and instability in Sraffian models. In L. L. Pasinetti (Ed.), Italian economic papers (Vol. 1, pp. 215–250). Oxford: Oxford University Press.

    Google Scholar 

  • Flaschel, P. (2010). Topics in classical micro- and macroeconomics. Elements of a critique of Neoricardian theory. Heidelberg: Springer.

    Book  Google Scholar 

  • Gnos, C., & Rochon, L.-P. (Eds.). (2008). The Keynesian multiplier. London: Routledge.

    Google Scholar 

  • Goldberg, G., & Neumann, M. (2003). Distribution of subdominant eigenvalues of matrices with random rows. Society for Industrial and Applied Mathematics Journal on Matrix Analysis and Applications, 24(3), 747–761.

    Google Scholar 

  • Goldberg, G., Okunev, P., Neumann, M., & Schneider, H. (2000). Distribution of subdominant eigenvalues of random matrices. Methodology and Computing in Applied Probability, 2(2), 137–151.

    Article  Google Scholar 

  • Goodwin, R. M. (1970). Elementary economics from the higher standpoint. Cambridge: Cambridge University Press.

    Google Scholar 

  • Gurgul, H., & Wójtowicz, T. (2015). On the economic interpretation of the Bródy conjecture. Economic Systems Research, 27(1), 122–131.

    Article  Google Scholar 

  • Halpern, L., & Molnár, G. (1997). Equilibrium and disequilibrium in a disaggregated classical model. In A. Simonovits & A. E. Stenge (Eds.), Prices, growth and cycles. Essays in honour of András Bródy (pp. 149–160). London: Macmillan.

    Chapter  Google Scholar 

  • Hartwig, J. (2004). Keynes’s multiplier in a two-sectoral framework. Review of Political Economy, 16(3), 309–334.

    Article  Google Scholar 

  • Horvath, M. (1998). Cyclicality and sectoral linkages: Aggregate fluctuations from independent sectoral shocks. Review of Economic Dynamics, 1(4), 781–808.

    Article  Google Scholar 

  • Horvath, M. (2000). Sectoral shocks and aggregate fluctuations. Journal of Monetary Economics, 45(1), 69–106.

    Article  Google Scholar 

  • Hua, L.-K. (1984). On the mathematical theory of globally optimal planned economic systems. Proceedings of the National Academy of Sciences of the United States of America, 81(20), 6549–6553.

    Article  Google Scholar 

  • Kaldor, N. (1955–1956). Alternative theories of distribution. The Review of Economic Studies, 23(2), 83–100.

    Google Scholar 

  • Katsinos, A., & Mariolis, T. (2012). Switch to devalued drachma and cost-push inflation: A simple input-output approach to the Greek case. Modern Economy, 3(2), 164–170.

    Article  Google Scholar 

  • Kurz, H. D. (1985). Effective demand in a ‘classical’ model of value and distribution: The multiplier in a Sraffian framework. The Manchester School, 53(2), 121–137.

    Article  Google Scholar 

  • Lange, O. (1967). The computer and the market. In C. H. Feinstein (Ed.), Socialism, capitalism and economic growth: Essays presented to Maurice Dobb (pp. 158–161). Cambridge: Cambridge University Press.

    Google Scholar 

  • Lange, O. (1970). Introduction to economic cybernetics. Oxford: Pergamon Press.

    Google Scholar 

  • Leontief, W., & Bródy, A. (1993). Money-flow computations. Economic Systems Research, 5(3), 225–233.

    Article  Google Scholar 

  • Long, J. B., Jr., & Plosser, C. I. (1983). Real business cycles. Journal of Political Economy, 91(1), 39–69.

    Article  Google Scholar 

  • Lucas, R. E. (1977). Understanding business cycles. Carnegie-Rochester Conference Series on Public Policy, 5(1), 7–29.

    Article  Google Scholar 

  • Mariolis, T. (2008a). The Sraffian multiplier: Theory and application. Internal Report of the Study Group on Sraffian Economics, 12 Nov 2008 (in Greek). Athens: Department of Public Administration, Panteion University.

    Google Scholar 

  • Mariolis, T. (2008b). Pure joint production, income distribution, employment and the exchange rate. Metroeconomica, 59(4), 656–665.

    Article  Google Scholar 

  • Mariolis, T. (2011). The wages-profits-growth relationships and the peculiar case of the Greek economy. In T. Mariolis (Ed.) (2016), Essays on the work of Dimitris Batsis (forthcoming).

    Google Scholar 

  • Mariolis, T. (2014). Modelling the devaluation of the Greek currency. Business Perspectives, 13(1), 1–5.

    Google Scholar 

  • Mariolis, T., & Soklis, G. (2014). The Sraffian multiplier for the Greek economy: Evidence from the supply and use table for the year 2010. MPRA Paper No 6253. http://mpra.ub.uni-muenchen.de/60253/. Accessed 12 Dec 2014.

  • Mariolis, T., & Tsoulfidis, L. (2014). On Bródy’s conjecture: Theory, facts and figures about instability of the US economy. Economic Systems Research, 26(2), 209–223.

    Article  Google Scholar 

  • Mariolis, T., Economidis, C., Stamatis, G., & Fousteris, N. (1997). Quantitative evaluation of the effects of devaluation on the cost of production (in Greek). Athens: Kritiki.

    Google Scholar 

  • Metcalfe, J. S., & Steedman, I. (1981). Some long-run theory of employment, income distribution and the exchange rate. The Manchester School, 49(1), 1–20.

    Article  Google Scholar 

  • Miller, R. E., & Blair, P. D. (2009). Input-output analysis: Foundations and extensions. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Molnár, G., & Simonovits, A. (1998). The subdominant eigenvalue of a large stochastic matrix. Economic Systems Research, 10(1), 79–82.

    Article  Google Scholar 

  • Samuelson, P. A. (1970). Law of conservation of the capital-output ratio. Proceedings of the National Academy of Sciences of the United States of America, 67(3), 1477–1479.

    Article  Google Scholar 

  • Steedman, I. (1984). Natural prices, differential profit rates and the classical competitive process. The Manchester School, 52(2), 123–140.

    Article  Google Scholar 

  • Steedman, I. (2000). Income distribution, foreign trade and the value-added vector. Economic Systems Research, 12(2), 221–230.

    Article  Google Scholar 

  • Sun, G.-Z. (2008). The first two eigenvalues of large random matrices and Brody’s hypothesis on the stability of large input-output systems. Economic Systems Research, 20(4), 429–432.

    Article  Google Scholar 

  • ten Raa, T. (2005). The economics of input-output analysis. Cambridge: Cambridge University Press.

    Google Scholar 

  • Thirlwall, A. P. (1979). The balance of payments constraint as an explanation of international growth rate differences. Banca Nazionale del Lavoro Quarterly Review, 32(128), 45–53.

    Google Scholar 

  • Thirlwall, A. P. (2011). Balance of payments constrained growth models: History and overview. PSL Quarterly Review, 64(259), 307–351.

    Google Scholar 

  • Trigg, Α. Β., & Philp, Β. (2008). Value magnitudes and the Kahn employment multiplier. Paper presented to Developing Quantitative Marxism, Bristol, 3–5 Apr 2008. http://carecon.org.uk/QM/Conference%202008/Papers/Trigg.pdf. Accessed 12 Dec 2014.

  • Tsaliki, P., & Tsoulfidis, L. (2015). Essays on political economy (in Greek). Thessaloniki: Tziolas.

    Google Scholar 

  • Tsoulfidis, L., & Tsaliki, P. (2014). Unproductive labour, capital accumulation and profitability crisis in the Greek economy. International Review of Applied Economics, 28(5), 562–585.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Appendices

Appendix 1: On the Sraffian Multiplier

The concept of the Sraffian multiplier, for a closed economy of single production with circulating capital, homogeneous labour and two types of income (wages and profits), was introduced by Kurz (1985).Footnote 10 This multiplier is an n × n matrix that depends on the (i) technical conditions of production, (ii) income distribution (and commodity prices), (iii) savings ratios out of wages and profits and (iv) consumption pattern s associated with the two types of income. Moreover, it includes, as special versions or limit cases, the usual Keynesian multiplier, the multipliers of the traditional input-output analysis and their Marxian versions.Footnote 11

Although in a quite different algebraic form, the Sraffian multiplier had been essentially introduced by Metcalfe and Steedman (1981) in a model with the following characteristics: open economy of single production with circulating capital, non-competitive imports , homogeneous labour and uniform rates of profits (and growth ), propensity to save and composition of consumption. Furthermore, Mariolis (2008b) (i) showed the mathematical equivalence between the Sraffian multiplier(s) derived from Kurz (1985) and Metcalfe and Steedman (1981) and (ii) extended the investigation of the latter to the case of pure joint production .

Assume that there are no non-competitive imports and that the price side of the system can be described by (see Sect. 2.2.3)

$$ \mathbf{p}\mathbf{B}=\mathbf{w}\widehat{\mathbf{l}}+\mathbf{p}\mathbf{A}\left[\mathbf{I}+\widehat{\overline{\mathbf{r}}}\right] $$
(6.1)

where w (w j > 0) denotes the 1 × n vector of money wage rates, \( \widehat{\mathbf{l}}\kern0.5em \left({l}_j>0\right) \) the n × n diagonal matrix of direct labour coefficients, \( \widehat{\overline{\mathbf{r}}} \) (r j > −1 and \( \widehat{\overline{\mathbf{r}}}\ne \mathbf{0} \)) the n × n diagonal matrix of the given (and constant) values of the sectoral profit rates and p is identified with e. Provided that [BA] is non-singular, Eq. 6.1 can be rewritten as

$$ \mathbf{p}=\mathbf{w}{\mathbf{V}}_{\mathbf{B}}+\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}} $$
(6.2)

where \( {\mathbf{V}}_{\mathbf{B}}\equiv \widehat{\mathbf{l}}{\left[\mathbf{B}-\mathbf{A}\right]}^{-1} \) denotes the matrix of additive labour values, and \( {\tilde{\mathbf{H}}}_{\mathbf{B}}\equiv \mathbf{A}\widehat{\overline{\mathbf{r}}}{\left[\mathbf{B}-\mathbf{A}\right]}^{-1} \).

Also assume that the quantity side of the economy can be described by (consider Sect. 2.2.1.4)

$$ \mathbf{B}{\mathbf{x}}^{\mathrm{T}}=\mathbf{A}{\mathbf{x}}^{\mathrm{T}}+{\mathbf{y}}^{\mathrm{T}} $$

or

$$ {\mathbf{x}}^{\mathrm{T}}={\left[\mathbf{B}-\mathbf{A}\right]}^{-1}{\mathbf{y}}^{\mathrm{T}} $$
(6.3)

and

$$ {\mathbf{y}}^{\mathrm{T}}={\mathbf{f}}_w^{\mathrm{T}}+{\mathbf{f}}_p^{\mathrm{T}}-\mathbf{I}{\mathbf{m}}^{\mathrm{T}}+{\mathbf{d}}^{\mathrm{T}} $$

or, setting \( \mathbf{I}{\mathbf{m}}^{\mathrm{T}}=\widehat{\mathbf{m}}\mathbf{B}{\mathbf{x}}^{\mathrm{T}} \),

$$ {\mathbf{y}}^{\mathrm{T}}={\mathbf{f}}_w^{\mathrm{T}}+{\mathbf{f}}_p^{\mathrm{T}}-\widehat{\mathbf{m}}\mathbf{B}{\mathbf{x}}^{\mathrm{T}}+{\mathbf{d}}^{\mathrm{T}} $$
(6.4)

where x T denotes the activity level vector, y T the vector of effective final demand, f T w the vector of consumption demand out of wages, f T p the vector of consumption demand out of profits, Im T the import demand vector, d T the autonomous demand vector (government expenditures , investments and exports ) and \( \widehat{\mathbf{m}} \) the diagonal matrix of imports per unit of gross output of each commodity.

If f T denotes the uniform consumption pattern (associated with the two types of income), s w denotes the saving ratio out of wages and s p denotes the saving ratio out of profits, where \( 0\le {s}_w,{s}_p\le 1 \) (and s w and s p are not both zero and not both unity), then the consumption demands out of wages and out of profits, in physical terms, amount to (see Eqs. 6.2 and 6.3, which imply that \( \widehat{\mathbf{l}}{\mathbf{x}}^{\mathrm{T}}={\mathbf{V}}_{\mathbf{B}}{\mathbf{y}}^{\mathrm{T}} \) and \( \mathbf{A}\widehat{\overline{\mathbf{r}}}{\mathbf{x}}^{\mathrm{T}}={\tilde{\mathbf{H}}}_{\mathbf{B}}{\mathbf{y}}^{\mathrm{T}} \))

$$ {\mathbf{f}}_w^{\mathrm{T}}=\left(1-{s}_w\right){\displaystyle \sum_{j=1}^n\left({w}_j{l}_j{x}_j\right)}{\left(\mathbf{p}{\mathbf{f}}^{\mathrm{T}}\right)}^{-1}{\mathbf{f}}^{\mathrm{T}}=\left(1-{s}_w\right)\left(\mathbf{w}\widehat{\mathbf{l}}{\mathbf{x}}^{\mathrm{T}}\right){\left(\mathbf{p}{\mathbf{f}}^{\mathrm{T}}\right)}^{-1}{\mathbf{f}}^{\mathrm{T}} $$

or

$$ {\mathbf{f}}_w^{\mathrm{T}}=\left(1-{s}_w\right)\left(\mathbf{w}{\mathbf{V}}_{\mathbf{B}}{\mathbf{y}}^{\mathrm{T}}\right){\left(\left(\mathbf{p}{\mathbf{f}}^{\mathrm{T}}\right)\right.}^{-1}{\mathbf{f}}^{\mathrm{T}} $$
(6.5)

and

$$ {\mathbf{f}}_p^{\mathrm{T}}=\left(1-{s}_p\right)\left(\mathbf{p}\mathbf{A}\widehat{\overline{\mathbf{r}}}{\mathbf{x}}^{\mathrm{T}}\right){\left(\mathbf{p}{\mathbf{f}}^{\mathrm{T}}\right)}^{-1}{\mathbf{f}}^{\mathrm{T}}=\left(1-{s}_p\right)\left(\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}{\mathbf{y}}^{\mathrm{T}}\right){\left(\mathbf{p}{\mathbf{f}}^{\mathrm{T}}\right)}^{-1}{\mathbf{f}}^{\mathrm{T}} $$
(6.6)

respectively.

Substituting Eqs. 6.5 and 6.6 into Eq. 6.4 leads to (take into account Eqs. 6.1 and 6.3 and that \( \left(\mathbf{w}{\mathbf{V}}_{\mathbf{B}}{\mathbf{y}}^{\mathrm{T}}\right){\mathbf{f}}^{\mathrm{T}}=\left({\mathbf{f}}^{\mathrm{T}}\mathbf{w}{\mathbf{V}}_{\mathbf{B}}\right){\mathbf{y}}^{\mathrm{T}} \), \( \left(\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}{\mathbf{y}}^{\mathrm{T}}\right){\mathbf{f}}^{\mathrm{T}}=\left({\mathbf{f}}^{\mathrm{T}}\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}\right){\mathbf{y}}^{\mathrm{T}} \)):

$$ {\mathbf{y}}^{\mathrm{T}}=\boldsymbol{\Lambda} {\mathbf{y}}^{\mathrm{T}}+{\mathbf{d}}^{\mathrm{T}} $$
(6.7)

where

$$ \boldsymbol{\Lambda} \equiv {\left(\mathbf{p}{\mathbf{f}}^{\mathrm{T}}\right)}^{-1}{\mathbf{f}}^{\mathrm{T}}\left[\mathbf{p}-\left({s}_w\mathbf{w}{\mathbf{V}}_B+{s}_p\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}\right)\right]-\widehat{\mathbf{m}}\mathbf{B}{\left[\mathbf{B}-\mathbf{A}\right]}^{-1} $$

Provided that \( \left[\mathbf{I}-\boldsymbol{\Lambda} \right] \) is non-singular, Eq. 6.7 can be uniquely solved for y T:

$$ {\mathbf{y}}^{\mathrm{T}}=\boldsymbol{\Pi} {\mathbf{d}}^{\mathrm{T}} $$

where \( \boldsymbol{\Pi} \equiv {\left[\mathbf{I}-\boldsymbol{\Lambda} \right]}^{-1} \) is the Sraffian multiplier linking autonomous demand to net output . Consequently, the change on the money value of net output, Δ i y (‘output multiplier’) induced by the increase of one unit of the autonomous demand for commodity i, is given by

$$ {\Delta}_y^i\equiv \mathbf{p}\boldsymbol{\Pi } {\mathbf{e}}_i^{\mathrm{T}} $$

If all the eigenvalues of Λ are less than 1 in modulus, then the dynamic multiplier process defined by (see Chipman 1950)

$$ {\mathbf{y}}_t^{\mathrm{T}}=\boldsymbol{\Lambda} {\mathbf{y}}_{t-1}^{\mathrm{T}}+\Delta {\mathbf{d}}^{\mathrm{T}},\kern0.5em t=1,2,\dots $$
(6.8)

is stable (also consider Sect. 2.2.2). In that case, the number \( - \log {\lambda}_{\max}\left[\boldsymbol{\Lambda} \right] \) is called the asymptotic rate of convergence (see, e.g. Berman and Plemmons 1994, Chap. 7) and provides a (rather) simple measure for the convergence rate of y T t to Πd T).

In the hypothetical (or heuristic) case where \( \widehat{\mathbf{m}}=\mathbf{0} \), Λ reduces to a rank-one matrix, i.e.

$$ {\boldsymbol{\Lambda}}_0\equiv {\left(\mathbf{p}{\mathbf{f}}^{\mathrm{T}}\right)}^{-1}{\mathbf{f}}^{\mathrm{T}}\left[\mathbf{p}-\left({s}_w\mathbf{w}{\mathbf{V}}_B+{s}_p\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}\right)\right] $$

the non-zero eigenvalue of which equals

$$ 1-{\left(\mathbf{p}{\mathbf{f}}^{\mathrm{T}}\right)}^{-1}\left({s}_w\mathbf{w}{\mathbf{V}}_B+{s}_p\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}\right){\mathbf{f}}^{\mathrm{T}} $$

or, invoking Eq. 6.2,

$$ 1-\left[{s}_w+\left({s}_p-{s}_w\right){\left(\mathbf{p}{\mathbf{f}}^{\mathrm{T}}\right)}^{-1}\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}{\mathbf{f}}^{\mathrm{T}}\right] $$

Thus, the Sraffian multiplier reduces to

$$ {\boldsymbol{\Pi}}_0\equiv {\left[\mathbf{I}-{\left(\mathbf{p}{\mathbf{f}}^{\mathrm{T}}\right)}^{-1}{\mathbf{f}}^{\mathrm{T}}\left[\mathbf{p}-\left({s}_w\mathbf{w}{\mathbf{V}}_{\mathbf{B}}+{s}_p\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}\right)\right]\right]}^{-1} $$

or, by applying the Sherman-Morrison formula ,

$$ {\boldsymbol{\Pi}}_0=\mathbf{I}+{\left(\left({s}_w\mathbf{w}{\mathbf{V}}_{\mathbf{B}}+{s}_p\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}\right){\mathbf{f}}^{\mathrm{T}}\right)}^{-1}{\mathbf{f}}^{\mathrm{T}}\left[\mathbf{p}-\left({s}_w\mathbf{w}{\mathbf{V}}_{\mathbf{B}}+{s}_p\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}\right)\right] $$
(6.9)

It then follows that (i) y T is not uniquely determined when

$$ \left({s}_w\mathbf{w}{\mathbf{V}}_{\mathbf{B}}+{s}_p\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}\right){\mathbf{f}}^{\mathrm{T}}=0 $$
(6.9a)

and (ii) one eigenvalue of Π 0 equals

$$ {\left[{s}_w+\left({s}_p-{s}_w\right){\left(\mathbf{p}{\mathbf{f}}^{\mathrm{T}}\right)}^{-1}\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}{\mathbf{f}}^{\mathrm{T}}\right]}^{-1} $$

while all the other eigenvalues equal 1. The former eigenvalue corresponds to a Kaldorian multiplier (see Kaldor 1955–1956) and could be conceived of as the system’s multiplier (associated with the case \( \widehat{\mathbf{m}}=\mathbf{0} \)). Furthermore, from Eqs. 6.2 and 6.9, it follows that when both wV B and \( \mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}} \) are semi-positive, (i) Π 0 is semi-positive, (ii) its diagonal elements are greater than or equal to 1 and (iii) its elements are non-increasing functions of s w and s p (as in the single production case; see Kurz 1985, pp. 133 and 135–136). For s w = s p = s (and p = e), each column of Π 0 sums to s −1. In the case of homogeneous labour and for s w = 0 and s p = 1, Π 0 reduces to a Marxian multiplier defined by Trigg and Philp (2008).

Finally, it seems that only little can be said, a priori, for the case where \( \widehat{\mathbf{m}}\ge \mathbf{0} \) (also consider Mariolis 2008b). For instance, the application of the previous analysis to the SUT of the Greek economy for the year 2010 (n = 63) gives the following resultsFootnote 12:

  1. (i)

    The matrix [BA]−1 (and, therefore, V B ) contains negative elements. Consequently, the system under consideration is not ‘all-productive’, and, therefore, it does not have the properties of a single-product system.

  2. (ii)

    The matrix \( \widehat{\mathbf{r}} \) contains one negative element. The matrix \( {\tilde{\mathbf{H}}}_{\mathbf{B}} \) contains negative elements, although some of its columns are positive. The vector \( \mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}\left(=\mathbf{e}{\tilde{\mathbf{H}}}_{\mathbf{B}}\right) \) contains one negative element, while all of its remaining elements are semi-positive and less than 1 (see Eq. 6.2). It then follows that there exist values of s w , s p , for which Π 0 is not semi-positive (see Eq. 6.9).

  3. (iii)

    For every value of s w , s p , it holds \( \left({s}_w\mathbf{w}{\mathbf{V}}_{\mathbf{B}}+{s}_p\mathbf{p}{\tilde{\mathbf{H}}}_{\mathbf{B}}\right){\mathbf{f}}^{\mathrm{T}}>0 \) (see Eq. 6.9a). Hence, the dynamic multiplier process defined by Λ 0 is stable, Π 0 is uniquely determined, and the eigenvalue of Π 0 that differs from 1 is approximately equal to \( {\left[{s}_w+\left({s}_p-{s}_w\right)0.655\right]}^{-1}\left(>1\right) \). For instance, for s w = 0 and s p = 1, Π 0 is semi-positive, its diagonal elements are in the range of 1–1.059, that eigenvalue is 1.527, and the arithmetic mean of the output multipliers, Δ i y , is 1.707. By contrast, the largest eigenvalue of Λ is −18.555 and its dominant eigenvalue ratio is 0.911 \( \left( rank\left[\boldsymbol{\Lambda} \right]=50\right) \). Matrix Π is not semi-positive, while its diagonal elements are all positive and in the range of 0.051–1.037. The dominant eigenvalue of Π is 1.219, the arithmetic mean of the moduli of its eigenvalues is 0.791, and that of Δ i y is 1.032. It can therefore be stated that the foreign sector plays a key role in the Sraffian multiplier for the Greek economy.

Appendix 2: Price Effects of Currency Devaluation

In what follows, let us assume that there are no non-competitive imports and that the price side of the system can be described by

$$ \mathbf{p}=\mathbf{p}\mathbf{A}+\mathbf{s} $$
(6.10)

or

$$ \mathbf{p}=\mathbf{p}{\mathbf{A}}^{\mathrm{d}}+\varepsilon {\mathbf{p}}^{\mathrm{m}}{\mathbf{A}}^{\mathrm{m}}+\mathbf{s} $$
(6.10a)

where p (= e) denotes the stationary price vector of domestically produced commodities; A d, A m the irreducible and primitive matrices of domestic and imported input-output coefficients , respectively; AA d + A m, with λ A1 < 1; ε the nominal exchange rate ; p m the given vector of foreign currency prices of the imported commodities; p = ε p m; and s (> 0) the vector of gross values added per unit activity level , which, in national accounts terms, equals the sum of consumption of fixed capital, s C, net taxes on production, s T, net operating surplus , s S, and compensation of employees , s E, i.e.

$$ \mathbf{s}\equiv {\mathbf{s}}_{\mathrm{C}}+{\mathbf{s}}_{\mathrm{T}}+{\mathbf{s}}_{\mathrm{S}}+{\mathbf{s}}_{\mathrm{E}} $$
(6.11)

By solving Eqs. 6.10 and 6.10a for p, we obtain

$$ \mathbf{p}=\mathbf{s}{\left[\mathbf{I}-\mathbf{A}\right]}^{-1}=\left(\varepsilon \mathbf{m}+\mathbf{s}\right){\left[\mathbf{I}-{\mathbf{A}}^{\mathrm{d}}\right]}^{-1} $$
(6.12)

where mp m A m. The price effect s of currency devaluation may be represented by the following dynamic version of Eq. 6.10a:

$$ {\mathbf{p}}_{t+1}={\mathbf{p}}_t{\mathbf{A}}^{\mathrm{d}}+{\varepsilon}_1\mathbf{m}+{\mathbf{s}}_t,\kern0.5em t=0,1,\dots $$
(6.13)

where \( {\varepsilon}_1\equiv \left(1+\overset{\frown }{\varepsilon}\right){\varepsilon}_0 \), \( \overset{\frown }{\varepsilon } \) denotes the devaluation rate and

$$ {\mathbf{p}}_0=\left({\varepsilon}_0\mathbf{m}+\mathbf{s}\right){\left[\mathbf{I}-{\mathbf{A}}^{\mathrm{d}}\right]}^{-1} $$

(see Eq. 6.12). Although there are alternative approaches for modelling the response of sectoral gross value added to currency devaluation , the choice between them should also take into account the input-output data availability. Thus, for the purposes of an indicative estimation, it may be postulated, for instance, that

$$ {\mathbf{s}}_t=\left({\mathbf{p}}_t{\mathbf{A}}^{\mathrm{d}}+{\varepsilon}_1\mathbf{m}+{\mathbf{p}}_t{\widehat{\mathbf{S}}}_{\mathrm{CT}}\right)\widehat{\boldsymbol{\upmu}}+{\mathbf{p}}_t{\widehat{\mathbf{S}}}_{\mathrm{CT}} $$

where (see Eq. 6.11) \( {\widehat{\mathbf{S}}}_{\mathrm{C}\mathrm{T}}\equiv \left[{\widehat{\mathbf{s}}}_{\mathrm{C}}+{\widehat{\mathbf{s}}}_{\mathrm{T}}\right]{\widehat{\mathbf{p}}}_0^{-1} \), \( \widehat{\boldsymbol{\upmu}}\equiv \left[{\widehat{\mathbf{s}}}_{\mathrm{S}}+{\widehat{\mathbf{s}}}_{\mathrm{E}}\right]{\widehat{\mathbf{c}}}_0^{-1} \) and

$$ {\mathbf{c}}_0\equiv {\mathbf{p}}_0{\mathbf{A}}^{\mathrm{d}}+{\varepsilon}_0\mathbf{m}+{\mathbf{s}}_{\mathrm{C}}+{\mathbf{s}}_{\mathrm{T}} $$

which imply that Eq. 6.13 becomes

$$ {\mathbf{p}}_{t+1}={\mathbf{p}}_t\boldsymbol{\Delta} +{\varepsilon}_1{\mathbf{m}}^{*} $$
(6.14)

where \( \boldsymbol{\Delta} \equiv \left[{\mathbf{A}}^{\mathrm{d}}+{\widehat{\mathbf{S}}}_{\mathrm{CT}}\right]\left[\mathbf{I}+\widehat{\boldsymbol{\upmu}}\right] \) and \( {\mathbf{m}}^{*}\equiv \mathbf{m}\left[\mathbf{I}+\widehat{\boldsymbol{\upmu}}\right] \). Then the solution of Eq. 6.14 is

$$ {\mathbf{p}}_t={\mathbf{p}}_0{\boldsymbol{\Delta}}^t+{\varepsilon}_1{\mathbf{m}}^{*}\left[{\boldsymbol{\Delta}}^{t-1}+{\boldsymbol{\Delta}}^{t-2}+\dots +\mathbf{I}\right] $$

and p t tends to \( {\varepsilon}_1{\mathbf{m}}^{*}{\left[\mathbf{I}-\boldsymbol{\Delta} \right]}^{-1}=\left(1+\widehat{\varepsilon}\right){\mathbf{p}}_0 \), since \( {\lambda}_{\boldsymbol{\Delta} 1}<1 \), while the price movement is governed by

$$ {\mathbf{m}}^{*}\left[{\boldsymbol{\Delta}}^{t-1}+{\boldsymbol{\Delta}}^{t-2}+\dots +\mathbf{I}\right] $$

which could be conceived of as the series of dated quantities of imported inputs.

In the extreme case where the gross values added are ‘insensitive’ to devaluation, i.e. s t = s, Δ should be replaced by A d, m * should be replaced by \( \mathbf{m}+{\varepsilon}_1^{-1}\mathbf{s} \), and p t tends to \( \left({\varepsilon}_1\mathbf{m}+\mathbf{s}\right){\left[\mathbf{I}-{\mathbf{A}}^{\mathrm{d}}\right]}^{-1}<\left(1+\widehat{\varepsilon}\right){\mathbf{p}}_0 \). Finally, in the ‘intermediate’ case where \( {\mathbf{s}}_t={\mathbf{p}}_t\widehat{\mathbf{s}}{\widehat{\mathbf{p}}}_0^{-1} \), Δ should be replaced by \( {\mathbf{A}}^{\mathrm{d}}+\widehat{\mathbf{s}}{\widehat{\mathbf{p}}}_0^{-1} \), m * should be replaced by m, and p t tends to \( \left(1+\widehat{\varepsilon}\right){\mathbf{p}}_0 \).Footnote 13

Empirical evidence from the SIOT of the Greek economy for the year 2005 shows that (Katsinos and Mariolis 2012):

  1. (i)

    The P-F eigenvalue of A d is 0.321 and the damping ratio is 1.290.

  2. (ii)

    The P-F eigenvalue of \( {\mathbf{A}}^{\mathrm{d}}+\widehat{\mathbf{s}}{\widehat{\mathbf{p}}}_0^{-1} \) is 0.949 and the damping ratio is 1.045.

  3. (iii)

    The P-F eigenvalue of \( \left[{\mathbf{A}}^{\mathrm{d}}+{\widehat{\mathbf{S}}}_{\mathrm{CT}}\right]\left[\mathbf{I}+\widehat{\boldsymbol{\upmu}}\right] \) is 0.893 and the damping ratio is 1.248.Footnote 14

  4. (iv)

    For \( \widehat{\varepsilon}=50\% \), the cost -inflation rate (as measured by the gross value of domestic production) at t = 1 is 9.3 % (is 5.3 %), and the arithmetic mean of commodity prices associated with matrix \( \left[{\mathbf{A}}^{\mathrm{d}}+{\widehat{\mathbf{S}}}_{\mathrm{CT}}\right]\left[\mathbf{I}+\widehat{\boldsymbol{\upmu}}\right] \) (with matrix \( {\mathbf{A}}^{\mathrm{d}}+\widehat{\mathbf{s}}{\widehat{\mathbf{p}}}_0^{-1} \)) reaches approximately 95 % of its asymptotic value at t = 14 (at t = 30).

These figures seem to be consistent with the finding that (other things constant) even ‘large’ devaluations would not imply great inflationary ‘pressures’.Footnote 15

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer Japan

About this chapter

Cite this chapter

Mariolis, T., Tsoulfidis, L. (2016). Bródy’s Stability and Disturbances. In: Modern Classical Economics and Reality. Evolutionary Economics and Social Complexity Science, vol 2. Springer, Tokyo. https://doi.org/10.1007/978-4-431-55004-4_6

Download citation

Publish with us

Policies and ethics