Skip to main content
Log in

The Poverty-Environment Trap in a Modified Romer Model with Dirtiness

  • Published:
Environmental Modeling & Assessment Aims and scope Submit manuscript

Abstract

The model presented in this paper describes an economy with endogenous technical change and polluting production techniques. The main question we want to address is whether the adoption of “dirty” production processes might lead to a sustainable unique steady state, or guarantee the emergence of multiple equilibria. The application of the original Bogdanov–Takens theorem allows us to characterize the regions of the parametric space where the model exhibits either a global indeterminate equilibrium or a poverty-environment trap.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. That is to say, two identically endowed economies (with the same initial stock of both physical and natural capital), with different abatement programs, may start at some point to follow completely different equilibrium paths towards the long-run steady state [14]. The rise of multiple equilibria can also generate a vicious poverty-environment trap, and the economy might not be able to escape a low-growth steady state associated to an unsustainable use of natural resources [19]. Local analysis may, thus, be misleading, since it refers to the vicinity of a stationary state, whereas the initial values of the state variables may not belong to such a neighborhood.

  2. Other papers conclude that a determinate saddle-path stable equilibrium can be reached, thanks to a growing level of new home-made inventories, without whom indeterminacy problems are very likely to emerge [7, 23, 30].

  3. The Bogdanov–Takens bifurcation theorem is largely used in mathematics, physics, and biology but has found limited application in economics. To the best of our knowledge, only [13] seems to have applied the methodology in a monetary model in a R 2 ambient space.

  4. Assume ρ is a time preference rate.

  5. This specific production function exhibits constant returns to scale at a disaggregate level because each firm takes z as given. On the contrary, a social planner can internalize this kind of externality, thus obtaining increasing returns.

  6. Negative values of γ are consistent with a slowdown of total factor productivity [25].

  7. Following [33], we consider the terms resource depletion and polluting emissions as interchangeable. The same applies for the concepts of stock of natural resources and environmental quality. Moreover, the broad definition of natural capital given in [17] is used in this paper.

  8. For example, only a fraction of the oil burnt in a production process (z) may serve to produce the final output (Y). The remaining is eventually pushed into the atmosphere as a resulting damaging emission load (P).

  9. Although technology is not directly linked to pollution in this model, we assume that each new inventory due to technological advance is cleaner than the previous one. This is consistent with the empirical evidence that developed economic systems also seek a less polluted environment to live in.

  10. Necessary condition for a maximum can be checked by studying the sign of all principal minors of the Hessian matrix for the control variables of the problem, whose determinant is formed by the following signs:

    $$\left\vert H\right\vert =\left\vert \begin{array}{ccc} - & 0 & 0 \\ 0 & - & 0 \\ 0 & 0 & - \end{array} \right\vert $$

    thus obtaining, |H 1| < 0, |H 2| > 0, |H 3| < 0.

  11. Since these bifurcation values are particularly useful for a subsequent numerical simulation of the poverty-environment trap, their complex mathematical solution is not shown here but is available upon request.

  12. The case where s = − 1 is very similar, so we leave it out of this demonstration.

  13. Which is topologically equivalent to the case where γ < 0 in the original model, .

References

  1. Aghion, P., & Howitt, P. (1998). Endogenous growth theory (2nd Ed.). Cambridge: MIT Press.

    Google Scholar 

  2. Antoci, A., Galeotti, M., Russu, P. (2011). Poverty trap and global indeterminacy in a growth model with open-access natural resources. Journal of Economic Theory, 146, 569–591.

    Article  Google Scholar 

  3. Arnold, L.G. (2000). Endogenous technological change: a note on stability. Economic Theory, 16, 219–226.

    Article  Google Scholar 

  4. Arnold, L.G. (2000). Stability of the market equilibrium in Romer’s model of endogenous technological change: a complete characterization. Journal of Macroeconomics, 22, 69–84.

    Article  Google Scholar 

  5. Asada, T., Semmler, W., Novak, A.J. (1998). Endogenous growth and the balanced growth equilibrium. Research in Economics, 52, 189–212.

    Article  Google Scholar 

  6. Ayong Le Kama, A., & Schubert, K. (2004). The consequences of an endogenous discounting depending on environmental quality. Environmental and Resource Economics, 28(1), 31–53.

    Article  Google Scholar 

  7. Bella, G. (2009). Polluting productions and sustainable economic growth: a local stability analysis. Journal of Service Science & Management, 2, 47–55.

    Article  Google Scholar 

  8. Benhabib, J., & Farmer, R. (1994). Indeterminacy and increasing returns. Journal of Economic Theory, 63, 19–46.

    Article  Google Scholar 

  9. Benhabib, J., & Farmer, R. (1996). Indeterminacy and sector-specific externalities. Journal of Monetary Economics, 17, 421–43.

    Article  Google Scholar 

  10. Benhabib, J., Meng, Q., Nishimura, K. (2000). Indeterminacy under constant returns to scale in multisector economies. Econometrica, 68, 1541–1548.

    Article  Google Scholar 

  11. Benhabib, J., & Perli, R. (1994). Uniqueness and indeterminacy: on the dynamics of endogenous growth. Journal of Economic Theory, 63, 113–142.

    Article  Google Scholar 

  12. Benhabib, J., Perli, R., Xie, D. (1994). Monopolistic competition, indeterminacy and growth. Ricerche Economiche, 48, 279–298.

    Article  Google Scholar 

  13. Benhabib, J., Schmitt-Grohé, S., Uribe, M. (2001). The perils of Taylor rules. Journal of Economic Theory, 96, 40–69.

    Article  Google Scholar 

  14. Boldrin, M., Been-Lon, C., Wang, P. (2004). Human capital, trade and public policy in rapidly growing economies. Cheltenham: Edward Elgar.

    Google Scholar 

  15. Bovenberg, A.L., & Smulders, S. (1996). Transitional impacts of environmental policy in an endogenous growth model. International Economic Review, 37, 861–893.

    Article  Google Scholar 

  16. Chamley, C. (1993). Externalities and dynamics in models of learning or doing. International Economic Review, 34, 583–609.

    Article  Google Scholar 

  17. Costanza, R., & Daly, H.E. (1992). Natural capital and sustainable development. Conservation Biology, 6, 37–46.

    Article  Google Scholar 

  18. Dinda, S. (2004). Environmental Kuznets curve hypothesis: a survey. Ecological Economics, 49, 431–55.

    Article  Google Scholar 

  19. Finco, M.V. (2009). Poverty-environment trap: a non linear probit model applied to rural areas in the North of Brazil. American-Eurasian Journal Agricultural Environmental and Science, 5(4), 533–539.

    Google Scholar 

  20. Freire, E., Gamero, E., Ponce, E., Franquelo, L.G. (1989). An algorithm for symbolic computation of center manifold. In P. Gianni (Ed.), Symbolic and algebraic computation, lecture notes in computer science (pp. 218–230). Berlin: Springer.

    Chapter  Google Scholar 

  21. Gamero, E., & Ponce, E. (1991). Normal forms for planar systems with nilpotent linear part. In R. Seydel, F.W. Schneider, T. Küpper, H. Troger (Eds.), Bifurcation and chaos: analysis, algorithms, applications, international series of numerical mathematics (pp. 123–127). Basel: Birkhäuser.

    Chapter  Google Scholar 

  22. Ghiglino, C., & Sorger, C. (2002). Poverty traps, indeterminacy, and the wealth distribution. Journal of Economic Theory, 105, 120–139.

    Article  Google Scholar 

  23. Gomez, M.A. (2005). Transitional dynamics in an endogenous growth model with physical capital, human capital and R&D. Studies in Nonlinear Dynamics & Econometrics, 9.

  24. Grimaud, A. (1999). Pollution permits and sustainable growth in a Schumpeterian model. Journal of Environmental Economics and Management, 38, 249–266.

    Article  Google Scholar 

  25. Kremer, M. (1993). Population growth and technological change: one million B.C. to 1990. The Quarterly Journal of Economics, 108(3), 681–716.

    Article  Google Scholar 

  26. Mattana, P., Nishimura, K., Shigoka, T. (2009). Homoclinic bifurcation and global indeterminacy of equilibrium in a two-sector endogenous growth model. International Journal of Economic Theory, 5, 1–23.

    Google Scholar 

  27. Musu, I. (1995). Transitional dynamics to optimal sustainable growth. Fondazione ENI Enrico Mattei Working Paper, 50, 95.

    Google Scholar 

  28. Nishimura, K., & Shigoka, T. (2006). Sunspots and Hopf bifurcations in continuous time endogenous growth models. International Journal of Economic Theory, 2, 199–216.

    Article  Google Scholar 

  29. Pittel, K. (2003). Sustainability and endogenous growth. Cheltenham: Edward Elgar.

    Google Scholar 

  30. Restrepo-Ochoa, S.I., & Vazquez, J. (2004). Cyclical features of the Uzawa–Lucas endogenous growth model. Economic Modelling, 21, 285–322.

    Article  Google Scholar 

  31. Romer, P.M. (1990). Endogenous technological change. Journal of Political Economy, 98, S71–S102.

    Article  Google Scholar 

  32. Slobodyan, S. (2007). Indeterminacy and stability in a modified Romer model. Journal of Macroeconomics, 29, 169–177.

    Article  Google Scholar 

  33. Smulders, S. (2000). Economic growth and environmental quality. In H. Folmer, L. Gabel (eds.), Principles of environmental and resource economics (pp. 602–664). Cheltenham: Edward Elgar.

    Google Scholar 

  34. Stern, D.I. (2004). The rise and fall of the environmental Kuznets curve. World Development, 32, 1419–1439.

    Article  Google Scholar 

  35. Wiggins, S. (1991). Introduction to applied nonlinear dynamical systems and chaos. New York: Springer.

    Google Scholar 

Download references

Acknowledgments

Financial support from Regione Autonoma della Sardegna (LR 7/10/2007 under the project “ Una rilettura del carattere dualistico dell’economia italiana alla luce dei nuovi schemi interpretativi emersi dal dibattito teorico ed empirico”) and from the Ministero dell’Istruzione, Università e Ricerca Scientifica, PRIN 2006 under the project “Poverty traps and multiple equilibria: a framework to interpret Mezzogiorno’s development”is gratefully acknowledged. Finally, I wish to thank two anonymous referees for their helpful comments, suggestions, and criticisms that have definitely helped to improve the quality of the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Giovanni Bella.

Appendix

Appendix

  1. 1.

    The arrow sufficiency theorem holds since the maximized Hamiltonian, evaluated along the optimal control variables, is concave in all the state variables, as we can simply check through the sign of the minors of the Hessian matrix, whose determinant implies the following:

    $$\left\vert H\right\vert =\left\vert \begin{array}{ccc} - & 0 & + \\ 0 & - & 0 \\ + & 0 & - \end{array} \right\vert $$

    and thus, |H 1| < 0, |H 2| > 0, and |H 3| < 0 iff A > 1, that is, the number of designs must necessarily be greater than one.

  2. 2.

    Transversality conditions for a free terminal state hold for all shadow prices and are given by

    $$\begin{array}{@{}rcl@{}} \lim\limits_{t\rightarrow \infty }\lambda Ke^{-\rho t} &=&\tilde{\lambda}e^{(1-2\sigma )gt}\tilde{K}e^{gt}e^{-\rho t}\\ &=&\tilde{\lambda}\tilde{K}e^{-(2\sigma g+\rho )t}=0 \\ \lim\limits_{t\rightarrow \infty }\mu Ee^{-\rho t} &=&\tilde{\mu}e^{(1-2\sigma )gt}\tilde{E}e^{gt}e^{-\rho t}\\ &=&\tilde{\mu}\tilde{E}e^{-(2\sigma g+\rho )t}=0 \\ \lim\limits_{t\rightarrow \infty }\vartheta Ae^{-\rho t} &=&\tilde{\vartheta}e^{(\rho -\varphi -\gamma )t}\tilde{A}e^{gt}e^{-\rho t}\\ &=&\tilde{\vartheta}\tilde{A}e^{(g-\varphi -\gamma )t}=0 \end{array} $$

    where \(\tilde{\lambda}, \tilde{\mu}, \tilde{\vartheta}\), and \(\tilde{K}, \tilde{E}, \tilde{A}\), are the shadow prices and the state values on the balanced growth path;

  3. 3.

    Moreover, for free time t, we need to show that \(\underset{ t\rightarrow \infty }{\lim }H=0 \), which is always verified due to convergence towards zero of both the discounted utility function, \(\underset{t\rightarrow \infty}{\lim }U(\cdot )e^{-\rho t}=0\), and all the multipliers, as proved above.

  4. 4.

    We need first to translate the fixed point associated to to the origin, by substituting \(\tilde{x}\equiv x-\bar{x}^{\ast }, \tilde{m}\equiv m-\bar{m}^{\ast }, \tilde{q}=q-\bar{q}^{\ast }\), and introduce the new trivial variables \(\mu =\gamma -\bar{\gamma}\) and \(\nu=\rho -\bar{\rho}\), such that

    $$\begin{array}{@{}rcl@{}} \overset{\cdot }{\tilde{x}} &=&-(\bar{\rho}+\nu )(\tilde{x}+\bar{x}^{\ast })\\ &&+\frac{(1-\alpha )(\bar{\gamma}+\mu )-\sigma (1+\bar{\gamma}+\mu )}{\sigma (1+\bar{\gamma}+\mu )}\\ &&\times(\tilde{m}+\bar{m}^{\ast })(\tilde{x}+\bar{x}^{\ast })+(\tilde{x}+\bar{x}^{\ast })^{2}\smallskip \\ \overset{\cdot }{\tilde{m}} &=&\frac{\theta +\alpha (\bar{\gamma}+\mu)(\varphi +\bar{\gamma}+\mu )}{(\bar{\gamma}+\mu )(1-\alpha )}(\tilde{m}+\bar{m}^{\ast})\\ &&-\frac{1+\alpha (\bar{\gamma}+\mu )}{1+\bar{\gamma}+\mu}(\tilde{m}+\bar{m}^{\ast })^{2}\\ &&+\frac{1+\bar{\gamma}+\mu }{(\bar{\gamma}+\mu )(1-\alpha )}(\tilde{m}+\bar{ m}^{\ast })(\tilde{q}+\bar{q}^{\ast })^{\omega}\\ \overset{\cdot }{\tilde{q}} &=&-\frac{(\bar{\rho}+\nu )}{\omega }(\tilde{q}+\bar{q}^{\ast })+\frac{(\tilde{m}+\bar{m}^{\ast })(\tilde{q}+\bar{q}^{\ast})^{1+\omega }}{\omega (\tilde{x}+\bar{x}^{\ast })}\\ &&+\frac{(1+\bar{\gamma}+\mu )}{\omega }(\tilde{q}+\bar{q}^{\ast })^{1+\omega} \end{array} $$
    (8)

A second-order Taylor expansion of the translated vector field (8) with respect to \((\tilde{x},\tilde{m},\tilde{q},\mu ,\nu) \) is therefore made, which leads to

$$\begin{array}{@{}rcl@{}} \left( \begin{array}{c} \overset{\cdot }{\tilde{x}} \\ \overset{\cdot }{\tilde{m}} \\ \overset{\cdot }{\tilde{q}} \end{array} \right) &=&\left( \begin{array}{ccc} \bar{x}^{\ast } & \frac{(1-\alpha )\bar{\gamma}-\sigma (1+\bar{\gamma})}{ \sigma (1+\bar{\gamma})}\bar{x}^{\ast } & 0 \\ 0 & \frac{\left( 1+\alpha \bar{\gamma}\right) \bar{m}^{\ast }}{1+\bar{\gamma} } & \frac{(1+\bar{\gamma})\bar{m}^{\ast }}{\bar{\gamma}(1-\alpha )} \\ -\frac{\bar{m}^{\ast }\bar{q}^{\ast }{}^{2}}{\bar{x}^{\ast }{}^{2}} & \frac{ \bar{q}^{\ast }{}^{2}}{\bar{x}^{\ast }} & \rho \end{array} \right) \left( \begin{array}{c} \tilde{x} \\ \tilde{m} \\ \tilde{q} \end{array} \right) \\ &&+\left( \begin{array}{ccc} \frac{(1-\alpha )\bar{m}^{\ast }}{(1+\bar{\gamma})^{2}}\mu -v & \frac{ (1-\alpha )\bar{x}^{\ast }}{(1+\bar{\gamma})^{2}}\mu & 0 \\ 0 & \left[ \frac{\alpha \bar{\gamma}^{2}-\theta }{(1-\alpha )\bar{\gamma}^{2} }+\frac{2(1-\alpha )}{(1+\bar{\gamma})^{2}}\bar{m}^{\ast }-\frac{1}{ (1-\alpha )(\bar{\gamma}+\mu )^{2}}\bar{q}^{\ast }\right] \mu & -\frac{\bar{m }^{\ast }}{(1-\alpha )\bar{\gamma}^{2}}\mu \\ 0 & 0 & 2\bar{q}^{\ast }\mu -v \end{array} \right) \left( \begin{array}{c} \tilde{x} \\ \tilde{m} \\ \tilde{q} \end{array} \right) \\ &&+\left( \begin{array}{c} \tilde{f}_{1}^{nl}(\tilde{x},\tilde{m},\tilde{q},\mu ,\nu ) \\ \tilde{f}_{2}^{nl}(\tilde{x},\tilde{m},\tilde{q},\mu ,\nu ) \\ \tilde{f}_{3}^{nl}(\tilde{x},\tilde{m},\tilde{q},\mu ,\nu ) \end{array} \right) \end{array} $$
(9)

where

$$\begin{array}{@{}rcl@{}} \tilde{f}_{1}^{nl}(\tilde{x},\tilde{m},\tilde{q},\mu ,\nu )&=&\tilde{x}^{2}- \frac{(1-\alpha )\bar{m}^{\ast }\bar{x}^{\ast }}{(1+\bar{\gamma})^{3}}\mu ^{2}-\frac{1+\alpha \bar{\gamma}}{1+\bar{\gamma}}\tilde{m}\tilde{x}\\ \tilde{f}_{2}^{nl}(\tilde{x},\tilde{m},\tilde{q},\mu ,\nu )&=&-\frac{1+\alpha \bar{\gamma}}{1+\bar{\gamma}}\tilde{m}^{2}+\frac{1+\bar{\gamma}}{\bar{\gamma} (1-\alpha )}\tilde{m}\tilde{q}+\left[ \frac{\theta \bar{m}^{\ast }}{ (1-\alpha )\bar{\gamma}^{3}}-\frac{(1-\alpha )\bar{m}^{\ast }{}^{2}}{(1+\bar{ \gamma})^{3}}-\frac{\bar{m}^{\ast }\bar{q}^{\ast }}{(1-\alpha )\bar{\gamma} ^{3}}\right] \mu ^{2}\\ \tilde{f}_{3}^{nl}(\tilde{x},\tilde{m},\tilde{q},\mu ,\nu )&=&\frac{\bar{m} ^{\ast }\bar{q}^{\ast }{}^{2}}{\bar{x}^{\ast }{}^{3}}\tilde{x}^{2}+\left[ \frac{\bar{m}^{\ast }}{\bar{x}^{\ast }}+(1+\bar{\gamma})\right] \tilde{q} ^{2}-\frac{\bar{q}^{\ast }{}^{2}}{\bar{x}^{\ast }{}^{2}}\tilde{x}\tilde{m}- \frac{2\bar{m}^{\ast }\bar{q}^{\ast }}{\bar{x}^{\ast }{}^{2}}\tilde{x}\tilde{ q}+\frac{2\bar{q}^{\ast }}{\bar{x}^{\ast }}\tilde{m}\tilde{q}\\ \end{array} $$

We ought to put the translated system (9) in the Jordan normal form characterized by a double-zero eigenvalue and a third eigenvalue given by the trace trJ . This computation is straightforward and needs

$$ \left( \begin{array}{c} \tilde{x} \\ \tilde{m} \\ \tilde{q} \end{array} \right) =\text{\textbf{T}}\left( \begin{array}{c} \tau _{1} \\ \tau _{2} \\ \tau _{3} \end{array} \right) $$
(10)

where T is a matrix of the eigenvectors

$$\text{\textbf{T}}=\left[ \begin{array}{ccc} \frac{j_{12}^{\ast }j_{23}^{\ast }}{j_{11}^{\ast }j_{22}^{\ast }} & \frac{ (j_{22}^{\ast })^{2}+j_{23}^{\ast }j_{32}^{\ast }}{j_{31}^{\ast }(j_{22}^{\ast })^{2}} & \frac{j_{12}^{\ast }j_{23}^{\ast }}{j_{23}^{\ast }(j_{22}^{\ast }+j_{33}^{\ast })} \\ -\frac{j_{23}^{\ast }}{j_{22}^{\ast }} & -\frac{j_{23}^{\ast }}{ (j_{22}^{\ast })^{2}} & 1 \\ 1 & 0 & \frac{j_{11}^{\ast }+j_{33}^{\ast }}{j_{23}^{\ast }} \end{array} \right] =\left[\begin{array}{ccc} u_{1} & v_{1} & z_{1} \\ u_{2} & v_{2} & 1 \\ 1 & 0 & z_{3} \end{array} \right] $$

which allows us to put the vector field (10) in the new coordinate change and the following normal form:

$$\begin{array}{@{}rcl@{}} \left( \begin{array}{c} \dot{\tau}_{1} \\ \dot{\tau}_{2} \\ \dot{\tau}_{3} \end{array} \right) &=&\left[ \begin{array}{ccc} 0 & 1 & 0 \\ 0 & 0 & 0 \\ 0 & 0 & \text{Tr}(\text{J}^{\mathbf{\ast }}(0)) \end{array} \right] \left( \begin{array}{c} \tau _{1} \\ \tau _{2} \\ \tau _{3} \end{array} \right)\\ &&+\mathbf{B}(\mu ,v)\left(\begin{array}{c} \tau _{1} \\ \tau _{2} \\ \tau _{3} \end{array} \right) +\left( \begin{array}{c} \bar{F}_{1}\left( \mathbf{\tau }\right) \\ \bar{F}_{2}\left( \mathbf{\tau }\right) \\ \bar{F}_{3}\left( \mathbf{\tau }\right) \end{array} \right) \end{array} $$
(11)

where \({\bf{\dot{\tau }}} = {{\rm{T}}^{{{\rm{ - 1}}}}}{J^{ * }}(0){\bf{T\tau }} + {{\bf{T}}^{{{\rm{ - 1}}}}}\tilde{f}_{i}^{{nl}}\left( {{\bf{T\tau }}} \right),{\bar{F}_{i}}\left( {\bf{\tau }} \right) = {{\bf{T}}^{{{\rm{ - 1}}}}}\tilde{f}_{i}^{{nl}}\left( {{\bf{T\tau }}} \right) = {{\bf{T}}^{{{\rm{ - 1}}}}}\tilde{f}_{i}^{{nl}}\left( {{\bf{Tw}}} \right)\) , and B(μ, v) is a matrix that vanishes at μ = 0, and v = 0:

$$ \text{\textbf{B}}(\mu ,v)=\text{\textbf{T}}^{\text{-1}}\left( \begin{array}{ccc} \frac{(1-\alpha )\bar{m}^{\ast }}{(1+\bar{\gamma})^{2}}\mu -v & \frac{ (1-\alpha )\bar{x}^{\ast }}{(1+\bar{\gamma})^{2}}\mu & 0 \\ 0 & \left[ \frac{\alpha \bar{\gamma}^{2}-\theta }{(1-\alpha )\bar{\gamma}^{2} }+\frac{2(1-\alpha )}{(1+\bar{\gamma})^{2}}\bar{m}^{\ast }-\frac{1}{ (1-\alpha )(\bar{\gamma}+\mu )^{2}}\bar{q}^{\ast }\right] \mu & -\frac{\bar{m }^{\ast }}{(1-\alpha )\bar{\gamma}^{2}}\mu \\ 0 & 0 & 2\bar{q}^{\ast }\mu -v \end{array} \right) \text{\textbf{T}} $$
(12)

The vector field (11) can be then suspended via a second-order center manifold reduction, which implies \(\dot{\tau}_{3}=0\), (see the mathematical procedure in [21]) and therefore becomes

$$\left( \begin{array}{c} \dot{\tau}_{1} \\ \dot{\tau}_{2} \end{array} \right) =\left( \begin{array}{cc} 0 & 1 \\ 0 & 0 \end{array} \right) \left( \begin{array}{c} \tau _{1} \\ \tau _{2} \end{array} \right) +\boldsymbol{A}(\mu ,v)\left( \begin{array}{c} \tau _{1} \\ \tau _{2} \end{array} \right) +\left( \begin{array}{c} F_{1} \\ F_{2} \end{array} \right) $$
(13)

where \(F_{1}+\mathcal{O}(w_{i}^{3})=\bar{F}_{1}\left( \boldsymbol{\tau }\right)\), and \(F_{2}+\mathcal{O}(w_{i}^{3})=\bar{F}_{2}\left( \boldsymbol{\tau }\right)\), and

$$\boldsymbol{A}(\mu ,\nu )=\frac{1}{D}\left( \begin{array}{cc} -v_{2}z_{3}A_{1}+v_{1}z_{3}B_{1}+(-v_{1}+v_{2}z_{1})C_{1} & -v_{2}z_{3}A_{2}+v_{1}z_{3}B_{2} \\ (u_{2}z_{3}-1)A_{1}+(z_{1}-u_{1}z_{3})B_{1}+(u_{1}-u_{2}z_{1})C_{1} & (u_{2}z_{3}-1)A_{2}+(z_{1}-u_{1}z_{3})B_{2} \end{array} \right) $$

with

$$\begin{array}{l}A_{1}=\left[ \frac{(1-\alpha )\bar{m}^{\ast }}{(1+\bar{\gamma})^{2}}\mu -v \right] u_{1}+\frac{(1-\alpha )\bar{x}^{\ast }}{(1+\bar{\gamma})^{2}}\mu u_{2} \\ A_{2}=\left[ \frac{(1-\alpha )\bar{m}^{\ast }}{(1+\bar{\gamma})^{2}}\mu -v\right] v_{1}+\frac{(1-\alpha )\bar{x}^{\ast }}{(1+\bar{\gamma})^{2}}\mu v_{2} \\ A_{3}=\left[ \frac{(1-\alpha )\bar{m}^{\ast }}{(1+\bar{\gamma})^{2}}\mu -v\right] z_{1}+\frac{(1-\alpha )\bar{x}^{\ast }}{(1+\bar{\gamma})^{2}}\mu \\ B_{1}=\left[ \frac{\alpha \bar{\gamma}^{2}-\theta }{(1-\alpha )\bar{\gamma}^{2}}+\frac{2(1-\alpha )}{(1+\bar{\gamma})^{2}}\bar{m}^{\ast }-\frac{1}{(1-\alpha )(\bar{\gamma}+\mu )^{2}}\bar{q}^{\ast }\right] \mu u_{2}-\frac{\bar{m}^{\ast }}{(1-\alpha )\bar{\gamma}^{2}}\mu \\ B_{2}=\left[ \frac{\alpha \bar{\gamma}^{2}-\theta }{(1-\alpha )\bar{\gamma}^{2}}+\frac{2(1-\alpha )}{(1+\bar{\gamma})^{2}}\bar{m}^{\ast }-\frac{1}{(1-\alpha )(\bar{\gamma}+\mu )^{2}}\bar{q}^{\ast }\right] \mu v_{2} \\ B_{3}=\left[ \frac{\alpha \bar{\gamma}^{2}-\theta }{(1-\alpha )\bar{\gamma}^{2}}+\frac{2(1-\alpha )}{(1+\bar{\gamma})^{2}}\bar{m}^{\ast }-\frac{1}{(1-\alpha )(\bar{\gamma}+\mu )^{2}}\bar{q}^{\ast }\right] \mu -\frac{\bar{m}^{\ast }}{(1-\alpha )\bar{\gamma}^{2}}\mu z_{3} \\C_{1}=2\bar{q}^{\ast }\mu -v \\C_{2}=0 \\ C_{3}=(2\bar{q}^{\ast }\mu -v)z_{3} \end{array} $$

Moreover, following Freire et al. [20] and Gamero et al. [21], we are able to reduce (13) to a topologically equivalent system:

$$\begin{array}{@{}rcl@{}} \dot{\tau}_{1}&=&\tau_{2}\\ \dot{\tau}_{2}&=&\varepsilon_{1}+\varepsilon_{2}\tau_{2}+\tau_{1}^{2}+s\tau_{1}\tau_{2} \end{array} $$
(14)

where ε 1 and ε 2 are the unfolding parameters of the versal deformation matrix, V \((\mu ,\nu )=\left ( \begin {array}{cc} 0 & 1 \\ \varepsilon _{1} & \varepsilon _{2} \end {array} \right ) \), that is

$$\begin{array}{rll} \varepsilon _{1}&=&\frac{\left[ -v_{2}z_{3}A_{2}+v_{1}z_{3}B_{2}\right] \left[ (u_{2}z_{3}-1)A_{1}+(z_{1}-u_{1}z_{3})B_{1}+(u_{1}-u_{2}z_{1})C_{1}\right] }{ D}- \\ && -\frac{\left[ -v_{2}z_{3}A_{1}+v_{1}z_{3}B_{1}+\left(-v_{1}+v_{2}z_{1}\right)C_{1} \right] \left[ (u_{2}z_{3}-1)A_{2}+(z_{1}-u_{1}z_{3})B_{2}\right] }{D} \end{array} $$
(15)
$$\begin{array}{rll} \varepsilon _{2}&=&\frac{-v_{2}z_{3}A_{1}+v_{1}z_{3}B_{1}+(-v_{1}+v_{2}z_{1})C_{1}}{D}+ \\ && +\frac{(u_{2}z_{3}-1)A_{2}+(z_{1}-u_{1}z_{3})B_{2}}{D} \end{array} $$
(16)

given that D = −v 1 + v 2 z 1u 1 v 2 z 3 + u 2 v 1 z 3.

Moreover, the stability of (15), i.e., ε 1 = 0, implies

$$ \xi _{1}\mu ^{2}+\xi _{2}v\mu +v_{1}v^{2}=0 $$
(17)

with

$$\begin{array}{l} \xi _{1}=b_{1}v_{2}z_{3}a_{1}-z_{3}\left[ b_{1}u_{2}-\frac{\bar{m}^{\ast }}{(1-\alpha )\bar{\gamma}^{2}}\right] c_{1}-2\bar{q}^{\ast }z_{1}b_{1}v_{2}+2 \bar{q}^{\ast }c_{1} \\ \xi _{2}=z_{3}\left[ b_{1}u_{2}-\frac{\bar{m}^{\ast }}{(1-\alpha )\bar{\gamma}^{2}}\right] v_{1}+z_{1}b_{1}v_{2}-b_{1}v_{2}z_{3}u_{1}-\left( 2\bar{q}^{\ast }v_{1}+c_{1}\right) \end{array}$$

where

$$\begin{array}{l} a_{1}=\left[ \frac{(1-\alpha )\bar{m}^{\ast }}{(1+\bar{\gamma})^{2}}u_{1}+\frac{(1-\alpha )\bar{x}^{\ast }}{(1+\bar{\gamma})^{2}}u_{2}\right] \\ b_{1}=\left[ \frac{\alpha \bar{\gamma}^{2}-\theta }{(1-\alpha )\bar{\gamma}^{2}}+\frac{2(1-\alpha )}{(1+\bar{\gamma})^{2}}\bar{m}^{\ast }-\frac{1}{(1-\alpha )\bar{\gamma}^{2}}\bar{q}^{\ast }\right] \\ c_{1}=\left[ \frac{(1-\alpha )\bar{m}^{\ast }}{(1+\bar{\gamma})^{2}}v_{1}+\frac{(1-\alpha )\bar{x}^{\ast }}{(1+\bar{\gamma})^{2}}v_{2}\right] \end{array} $$

Solution to (17) easily gives µ = Φv, where \(\Phi =\frac{\sqrt{\xi _{2}^{2}-4v_{1}\xi _{1}}-\xi _{2}}{2\xi _{1}}.\).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Bella, G. The Poverty-Environment Trap in a Modified Romer Model with Dirtiness. Environ Model Assess 18, 629–638 (2013). https://doi.org/10.1007/s10666-013-9372-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10666-013-9372-4

Keywords

JEL Classifications

Navigation