Skip to main content
Log in

Is Recycling a Threat or an Opportunity for the Extractor of an Exhaustible Resource?

  • Published:
Environmental and Resource Economics Aims and scope Submit manuscript

Abstract

We examine the best extraction strategies for the provider of an exhaustible resource that can be recycled. In a two-period model of resource extraction, the extractor faces prospective entry by a recycler that incurs a fixed cost to produce a perfect substitute of the virgin resource. Its entry is an opportunity or a threat for the extractor, depending on whether it maximizes social welfare or its own revenue. Our results highlight how prospective recycling modifies the Hotelling rule. We characterize various entry possibilities. The benevolent extractor may accommodate or promote recycling, while the self-interested extractor may accommodate or deter recycling.

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

Similar content being viewed by others

Notes

  1. There are various methods available to recover phosphorus, such as ploughing crop residues back into the soil, composting food waste from households, using human and animal excreta, and so on.

  2. “Around 41% of phosphorus from sewage sludge across the European Union is currently recovered and reused in agriculture” , from the European Commission’s expert seminar on the sustainability of phosphorus resources (2011, http://ec.europa.eu/environment/natres/pdf/conclusions_17_02_2011.pdf.). Even now, according to Ensink et al. (2004), more than 25% of urban vegetables grown in Pakistan are being fertilized with municipal wastewater.

  3. In 1989, about 28% of the total aluminum supply in the United States came from recycled aluminum (see http://www.epa.gov/osw/nonhaz/municipal/pubs/sw90077a.pdf).

  4. In 1945, Alcoa was judged to enjoy a strong monopoly position, which was supported rather than threatened by competition from secondary aluminum produced by recycling scrap aluminum. Swan (1980) provides empirical evidence that the price charged by Alcoa is only slightly below the pure monopoly price but is well above the purely competitive price. The question of whether Alcoa had maintained its monopoly position by strategically controlling the supply of scrap aluminum that was ultimately available to secondary producers has been debated at length in the economic literature. Grant (1999) provides a useful survey of this debate.

  5. Regarding phosphorus, for instance, the recycler may be viewed as the group of developed countries with phosphorus-saturated soils and advanced wastewater treatment technologies (see Weikard and Seyhan 2009).

  6. See Martin 1982 for aluminum scrap recycling and Weikard and Seyhan (2009) for phosphorus recycling from sewage sludge.

  7. \( P_{t}^{\prime }(.)\) denotes the derivative of \(P_{t}(x)\) with respect to x. Throughout the text, the superscripts \(^{\prime }\) and \(^{\prime \prime }\) will be used to represent, respectively, the first and second derivatives of a function of a single variable.

  8. A function \(P_{t}(.)\) is log-concave if \(P_{t}^{^{\prime \prime }}(.)P_{t}(.)-P_{t}^{\prime }(.)^{2}<0\). This assumption is very general: LogP is concave if P is concave, linear or \(P(q)=Aq^{\gamma -1}\) with \( 0<\gamma <1\) so that \(1/(1-\gamma )\) is the elasticity of demand. Most of the commonly used demand functions are, in fact, log-concave. The limiting case is \(P(q)=Ae^{-q}\), which is strictly convex and log-linear (hence log-concave).

  9. For a function with several variables, the partial derivative with respect to a variable is denoted by that variable’s being written as a subscript.

  10. We need this restriction because \({\widehat{q}}=\frac{\left( 1+{\widehat{\mu }} \right) ^{2}(a-c)-s{\widehat{\mu }}^{2}}{1+2{\widehat{\mu }}-{\widehat{\mu }}^{2}}\) and \({\widehat{r}}=\frac{\left( 1+{\widehat{\mu }}\right) (2(a-c)-s)}{1+2 {\widehat{\mu }}-{\widehat{\mu }}^{2}}\).

  11. In other terms, in the case \(\lambda =1\), we will restrict attention to the pure strategy in which the recycler chooses to stay out with probability 1 when the initial extraction is \({\widetilde{q}}\).

  12. In other terms, in the case \(\lambda =0\), we will restrict attention to the pure strategy in which the recycler chooses to enter with probability 1 when the initial extraction is \({\widetilde{q}}\).

  13. Otherwise, there can be no equilibrium at \(E^{\prime }\) so long as the extractor thinks that the recycler stays out with a positive probability.

  14. See Stiglitz (1976), for instance.

  15. Otherwise, there can be no equilibrium at H so long as the extractor thinks that there is a positive probability of entry.

  16. Let \(p_{2}^{N}\) and \(r^{N}\) respectively denote the market clearing price and the recycler’s output in the second period Nash equilibrium with no resource exaustion that would result from competition between the extractor and the recycler. Within this specific framework, this equilibrium outcome is characterized by \(q^{N}=\frac{a+c}{3},r^{N}=\frac{a-2c}{3}\) and \( p_{2}^{N}=\frac{a+c}{3}\).

References

  • André FJ, Cerdá E (2006) On the dynamics of recycling and natural resources. Environ Resour Econ 33(2):199–221

    Article  Google Scholar 

  • Bain JS (1956) Barriers to new competition. Harvard University Press, Cambridge

    Book  Google Scholar 

  • Baksi S, Long NV (2009) Endogenous consumer participation and the recycling problem. Aust Econ Pap 48(4):281–295

    Article  Google Scholar 

  • Baumol WJ, Bradford DF (1970) Optimal departures from marginal cost pricing. Am Econ Rev 60(3):265–283

    Google Scholar 

  • Boiteux M (1956) Sur la Gestion des Monopoles Publics Astreints à l’Equilibre Budgétaire. Econometrica 24(1):22–40; translated as On the management of public monopolies subject to budgetary constraints. J Econ Theory 3:219–240 (1971)

  • Cordell D, Drangert J-O, White S (2009) The story of phosphorus: global food security and food for thought. Glob Environ Change 19:292–305

    Article  Google Scholar 

  • Ensink JHJ, Mahmood T, van der Hoek W, Raschid-Sally L, Amerasinghe FP (2004) A nationwide assessment of wastewater use in Pakistan: an obscure activity or a vitally important one? Water Policy 6:197–206

    Article  Google Scholar 

  • Friedman M (1967) Price theory. Aldine Publishing Co., Chicago

    Google Scholar 

  • Gaudet G (2007) Natural resource economics under the rule of hotelling. Can J Econ 40:1033–1059

    Article  Google Scholar 

  • Gaudet G, Van Long N (2003) Recycling redux: a Nash–Cournot approach. Jpn Econ Rev 54(4):409–419

    Article  Google Scholar 

  • Gaudet G, Lasserre P, Van Long N (1995) Optimal resource royalties with unknown and temporally independent extraction cost structures. Int Econ Rev 36(3):715–749

    Article  Google Scholar 

  • Gilbert RJ, Goldman SM (1978) Potential competition and the monopoly price of an exhaustible resource. J Econ Theory 17(2):319–331

    Article  Google Scholar 

  • Grant D (1999) Recycling and market power: a more general model and re-evaluation of the evidence. Int J Ind Organ 17:59–80

    Article  Google Scholar 

  • Hollander A, Lasserre P (1988) Monopoly and the pre-emption of competitive recycling. Int J Ind Organ 6:489–497

    Article  Google Scholar 

  • Hotelling H (1931) The economics of exhaustible resources. J Polit Econ 39(2):137–175

    Article  Google Scholar 

  • Livingston D, Colebatch H, Ashbolt N (2005) Sustainable water paradigm shift: does changing discourse mean change in organisation? Working paper, School of Civil and Environmental Engineering, University of New South Wales, Sydney, Australia

  • Martin R (1982) Monopoly power and the recycling of raw materials. J Ind Econ 30(4):405–419

    Article  Google Scholar 

  • Maskin E, Tirole J (1987) A theory of dynamic oligopoly, III. Eur Econ Rev 31:947–968

    Article  Google Scholar 

  • Peterson FM, Fisher AC (1977) The exploitation of extractive resources: a survey. Econ J 87(348):681–721

    Article  Google Scholar 

  • Ramsey FP (1972) A contribution to the theory of taxation. Econ J 37(145):47–61

    Article  Google Scholar 

  • Sartorius C (2011) Lead market potential for phosphorus recycling technologies in Germany. In: Rongping M, Meckl R (eds) Innovation for green growth. Science Press, Beijing, pp 275–296

    Google Scholar 

  • Seyhan D, Weikard H-P, Ierland EV (2012) An economic model of long-term phosphorus extraction and recycling. Resour Conserv Recycl 61:103–108

    Article  Google Scholar 

  • Slade ME, Thille H (2009) Whither hotelling: tests of the theory of exhaustible resources. Ann Rev Resour Econ 1:239–260

    Article  Google Scholar 

  • Stiglitz J (1976) Monopoly and the rate of extraction of exhaustible resources. Am Econ Rev 66(4):655–661

    Google Scholar 

  • Suslow V (1986) Estimating monopoly behavior with competitive recycling: an application to Alcoa. RAND J Econ 17(3):389–403

    Article  Google Scholar 

  • Swan PL (1980) The influence of recycling on monopoly power. J Polit Econ 88(1):76–99

    Article  Google Scholar 

  • Tullock G (1979) Monopoly and the rate of extraction of exhaustible resources. Am Econ Rev 69(1):231–233

    Google Scholar 

  • Wan R, Boyce JR (2014) Non-renewable resource Stackelberg games. Resour Energy Econ 37:102–121

    Article  Google Scholar 

  • Weikard HP, Seyhan D (2009) Distribution of phosphorus resources between rich and poor countries: the effect of recycling. Ecol Econ 68:1749–1755

    Article  Google Scholar 

  • Weinstein MC, Zeckhauser RJ (1974) Use patterns for depletable and recycleable resources, symposium on the economics of exhaustible resources. Rev Econ Stud 41:67–88

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Philippe Mahenc.

Additional information

We are grateful to a co-editor and two anonymous referees for their insightful comments and suggestions. This research received funding from the French Agence Nationale de la Recherche within the framework of the project “GREENGO – New Tools for Environmental Governance: the role of NGOs” (ANR-15-CE05-0008). Financial support from the program “Investissements d’Avenir” of the French government (ANR-10-LABX-14-01) is gratefully acknowledged, too.

Appendix

Appendix

1.1 Appendix 1: Proof of Lemma 1

In the first step, we show that the solution \(\widehat{r }(q,\mu )\) of \(W_{r}^{n}(q,r)+\mu \pi _{r}(q,r)=0\) is a local maximum.

Let the Lagrangian be \(L(q,r)=W^{n}(q,r)+\mu \left[ P_{2}(s-q+r)r-c(r)-F \right] \). If \(L_{r}(q,r)=0\) has an interior solution \({\widehat{r}}(q,\mu )\), then we have

$$\begin{aligned} {\widehat{r}}(q,\mu )=-\frac{\mu +\delta }{\mu }\frac{P_{2}(s-q+r)-c^{\prime }(r)}{P_{2}^{\prime }(s-q+r)}. \end{aligned}$$
(39)

We now check that the second-order condition for \({\widehat{r}}(q,\mu )\) to be a local maximum of L(qr) is satisfied. By substituting \({\widehat{r}} (q,\mu )\) for r in

$$\begin{aligned} L_{rr}\left( q,r\right) =&\delta \left[ P_{2}^{\prime }(s-q+r)-c^{\prime \prime }(r)\right] +\mu \left[ 2P_{2}^{\prime }(s-q+r)\right. \nonumber \\&\left. +P_{2}^{\prime \prime }(s-q+r)r-c^{\prime \prime }(r)\right] , \end{aligned}$$
(40)

we obtain

$$\begin{aligned}&L_{rr}(q,{\widehat{r}}(q,\mu )) \nonumber \\&\quad =(\delta +\mu )\left[ P_{2}^{\prime }(s-q+{\widehat{r}}(q,\mu ))-c^{\prime \prime }({\widehat{r}}(q,\mu ))\right. \nonumber \\&\qquad \left. -\,\frac{P_{2}(s-q+{\widehat{r}}(q,\mu ))-c^{\prime }({\widehat{r}}(q,\mu ))}{P_{2}^{\prime }(s-q+{\widehat{r}}(q,\mu ))} P_{2}^{\prime \prime }(s-q+{\widehat{r}}(q,\mu ))\right] \nonumber \\&\qquad +\,\mu P_{2}^{\prime }(s-q+{\widehat{r}}(q,\mu )). \end{aligned}$$
(41)

Given that \(P_{2}(.)\) is log-concave by assumption \(\left( A2\right) \), we know that \(P_{2}^{\prime \prime }(.)<\frac{P_{2}^{\prime }(.)^{2}}{P_{2}(.)}\). Thus, we have

$$\begin{aligned} L_{rr}(q,{\widehat{r}}(q,\mu ))<\mu P_{2}^{\prime }(.)-\left( \delta +\mu \right) c^{\prime \prime }(.)+\frac{\left( \delta +\mu \right) c^{\prime }(.) }{P_{2}(.)}P_{2}^{\prime }(.). \end{aligned}$$
(42)

Under assumption \(\left( A1\right) \), the right-hand side of (42) is negative, and so \({\widehat{r}}(q,\mu )\) is a local maximum.

In the second step, we show that \(0<{\widehat{r}}_{q}(q,\mu )<1\).

Differentiating \(L_{r}(q,{\widehat{r}}(q,\mu ))=0\) yields the slope of \( {\widehat{r}}_{q}(q,\mu )\) with respect to q:

$$\begin{aligned} {\widehat{r}}_{q}(q,\mu )=-\frac{L_{rq}}{L_{rr}},=\frac{\delta P_{2}^{\prime }(.)+\mu \left[ P_{2}^{\prime }(.)+P_{2}^{\prime \prime }(.)r\right] }{ \delta \left[ P_{2}^{\prime }(.)-c^{\prime \prime }(.)\right] +\mu \left[ 2P_{2}^{\prime }(.)+P_{2}^{\prime \prime }(.)r-c^{\prime \prime }(.)\right] } . \end{aligned}$$
(43)

where \(-L_{rq}=\delta P_{2}^{\prime }(.)+\mu \left[ P_{2}^{\prime }(.)+P_{2}^{\prime \prime }(.)r\right] \). As \(L_{rr}<0\), we have sign \( \left( {\widehat{r}}_{q}\right) =\) sign \(\left( L_{rq}\right) \). After substitution of \({\widehat{r}}(q,\mu )\) for r, we obtain

$$\begin{aligned} -L_{rq}=(\delta +\mu )\left[ P_{2}^{\prime }(.)-\left( P_{2}(.)-c^{\prime }(.)\right) P_{2}^{\prime \prime }(.)/P_{2}^{\prime }(.)\right] . \end{aligned}$$
(44)

Again using the log-concavity of \(P_{2}(.)\), we have

$$\begin{aligned} -L_{rq}<(\delta +\mu )\left[ P_{2}^{\prime }(.)-\left( P_{2}(.)-c^{\prime }(.)\right) \frac{P_{2}^{\prime }(.)}{P_{2}(.)}\right] . \end{aligned}$$
(45)

The right-hand side of (45) reduces to \(\left( \delta +\mu \right) c^{\prime }(.)\frac{P_{2}^{\prime }(.)}{P_{2}(.)}<0\). It follows that \(-L_{rq}<0\) and, therefore, \({\widehat{r}}_{q}(q,\mu )>0\).

Further calculations yield

$$\begin{aligned} {\widehat{r}}_{q}(q,\mu )-1=\frac{\delta c^{\prime \prime }(.)+\mu \left[ -P_{2}^{\prime }(.)+c^{\prime \prime }(.)\right] }{L_{rr}}. \end{aligned}$$
(46)

The numerator of the fraction given on the right-hand side of (46) is strictly positive, whereas the denominator \(L_{rr}\) is negative. We can conclude that \({\widehat{r}}_{q}(q,\mu )-1<0\).

In the third step, we show that \({\widehat{r}}(q,\mu )\) is an interior solution under assumption \(\left( A3\right) \), which falls short of q.

Given that recycling cannot emerge ex nihilo, a minimum amount \( {{{\underline{q}}}}>0\) of the virgin resource is needed to produce the first unit of recycled output; that is, \({\widehat{r}}({{{\underline{q}}}},\mu )=0\). As \( {{{\underline{q}}}}>0\), we have \({\widehat{r}}(0,\mu )<{\widehat{r}}({{{\underline{q}}}},\mu )=0\) since \({\widehat{r}}\left( q,\mu \right) \) is upward sloping with respect to q. This inequality is equivalent to \(L_{r}(0,0)=-\left( \mu +\delta \right) \left[ P_{2}(s)-c^{\prime }(0)\right] \le 0\), which holds under assumption \(\left( A3\right) \). Moreover, \(L_{r}(s,0)=\left( \mu +\delta \right) \left[ P_{2}(0)-c^{\prime }(0)\right] \) is positive under assumption \(\left( A3\right) \). Thus, \({\widehat{r}}({{{\underline{q}}}},\mu )=0< {\widehat{r}}(s,\mu )\), and so \({{{\underline{q}}}}<s\). Finally, for all q higher than \({{{\underline{q}}}}\), we have \({\widehat{r}}(q,\mu )-q<{\widehat{r}}({{\underline{q }}},\mu )-{{{\underline{q}}}}\) since \({\widehat{r}}_{q}(q,\mu )-1<0\), and thus \( {\widehat{r}}(q,\mu )<q\): the virgin resource is never fully recycled.

1.2 Appendix 2: Proof of Proposition 1

The social planner maximizes \(W^{n}(q,r)\) under the break-even constraint (2).

Assuming an interior solution, the first-order conditions for this problem are

$$\begin{aligned} P_{1}(q)-\delta P_{2}(s-q+r)= & {} \mu P_{2}^{\prime }(s-q+r)r, \end{aligned}$$
(47)
$$\begin{aligned} P_{2}(s-q+r)-c^{\prime }(r)= & {} -\frac{\mu }{\left( \delta +\mu \right) } P_{2}^{\prime }(s-q+r)r, \end{aligned}$$
(48)
$$\begin{aligned} \mu \left[ P_{2}(s-q+r)r-c(r)-F\right]= & {} 0. \end{aligned}$$
(49)

If \(\mu =0\), (48) immediately shows that the recycling output must be sold at marginal cost. If \(\mu \) is positive, then the constraint (2) is binding and the social planner must use the proceeds from sales of the recycled product to finance the recycling costs. Using \(\varepsilon \), the first-order condition (48) yields the standard Ramsey formula

$$\begin{aligned} \frac{P_{2}(s-q+r)-c^{\prime }(r)}{P_{2}(s-q+r)}=\frac{\mu }{\left( \delta +\mu \right) }\frac{1}{\varepsilon }. \end{aligned}$$
(50)

Note that (50) can be rewritten as follows:

$$\begin{aligned} p_{2}\left( 1-\frac{\mu }{\left( \delta +\mu \right) \varepsilon }\right) =c^{\prime }(r). \end{aligned}$$
(51)

Thus, we must have \(\varepsilon >\frac{\mu }{\delta +\mu }\) for an interior solution. If this were not the case, the left-hand side of (51) would be negative and, hence, could not be equal to marginal cost.

Furthermore, setting \(\varepsilon \) in (47) yields:

$$\begin{aligned} \frac{p_{2}}{p_{1}}=\frac{1}{\delta }+\frac{\mu }{\delta \varepsilon }\frac{ p_{2}}{p_{1}}. \end{aligned}$$
(52)

By substituting \(\delta =\frac{1}{1+\rho }\) into (52), we obtain the variant of the Hotelling rule (8). Moreover, we can write (52) as

$$\begin{aligned} \frac{p_{2}}{p_{1}}\left( 1-\frac{\mu }{\delta \varepsilon }\right) =\frac{1 }{\delta }, \end{aligned}$$
(53)

which shows that \(\varepsilon \) must be higher than \(\frac{\mu }{\delta }\) to get an interior solution. If so, then the condition \(\varepsilon >\frac{ \mu }{\delta +\mu }\) needed for the existence of an interior solution of Eq. (51) is also satisfied.

1.3 Appendix 3: Proof of Corollary 2

With linear demand and cost functions, assumption \(\left( A4\right) \) can be written asFootnote 16

$$\begin{aligned} s<\frac{a+c}{3}. \end{aligned}$$
(54)

By substituting \(r^{*}(q)\) for r in (24), we obtain the reduced-form function

$$\begin{aligned} {\mathbf {V}}(q)=\left\{ \begin{array}{ll} aq-q^{2}/2+a(a-c+s-q)/2-(a-c+s-q)^{2}/8&{}\quad \text { if }{\widetilde{q}}\le q\le s,\\ \\ -q^{2}+sq+as-s^{2}/2&{}\quad \text { otherwise.} \end{array} \right. \end{aligned}$$
(55)

Assumption (29) boils down to \(W\left( q_{0},0\right) <{\mathbf {V}}(s)\), which requires that s belongs to the interval \(\left( a-c,\sqrt{\frac{\left( a-c\right) \left( 3a+c\right) }{2}} \right) \).

Substituting (28) into (25 ) yields

$$\begin{aligned} r\left( q_{a}\right) =\frac{4a-3c-2s}{5}. \end{aligned}$$
(56)

From (28), we see that \(q_{a}\le s\) provided that \(c\ge 3a-4s\). Otherwise, we have \({\widetilde{q}}\le s<q_{a}\) when \(F\le {\overline{F}}\), and thus \({\mathbf {V}}(q)\) achieves a global maximum at s from (29). Moreover, \(r\left( q_{a}\right) \ge 0\) requires that \(c\le \frac{2}{3}\left( 2a-s\right) \), which holds under (29). It turns out that \( F_{a}<{\overline{F}}\) when \(c\ge 3a-4s\) holds. One can also check that, under assumption \(\left( A4\right) \) and \(c<a\), we have \(q_{0}<q_{a}\).

There are three possibilities:

  1. (i)

    \({\widetilde{q}}<q_{0}<q_{a}<s\). The inequality \({\widetilde{q}} <q_{0} \) requires that \(F<F_{d}\), where \(F_{d}=\frac{(2\left( a-c\right) -s)^{2}}{16}\). Under (29), we have \( F_{d}<F_{a}\) because this requires that \(c<\frac{26a-13s}{22}\) and we have \( \sqrt{\frac{\left( a-c\right) \left( 3a+c\right) }{2}}<\frac{26a-13s}{22}\). Then, \({\mathbf {V}}(q) \) achieves a unique maximum at \(q_{a}\).

  2. (ii)

    \(q_{0}\le {\widetilde{q}}\le q_{a}<s\). This requires that \( F_{d}\le F\le F_{a}\). Under (29), \(\mathbf {V }(q)\) achieves a global maximum at \(q_{a}\) because \(W\left( q_{0},0\right)< {\mathbf {V}}(s)<{\mathbf {V}}(q_{a})\).

  3. (iii)

    \(q_{0}\le q_{a}<{\widetilde{q}}<s\). This requires that \(F_{a}<F< {\overline{F}}\). Under (29), \({\mathbf {V}}(q)\) achieves a global maximum at \({\widetilde{q}}\) because \(W\left( q_{0},0\right)<{\mathbf {V}}(s)<{\mathbf {V}}({\widetilde{q}})\).

1.4 Appendix 4: Proof of Corollary 3

By substituting \(r^{*}(q)\) for r in (34) within the framework of linear demand and cost functions, we obtain the reduced-form function

$$\begin{aligned} {\mathbf {V}}(q)=\left\{ \begin{array}{ll} (a-q)q+(a-s+q+c)(s-q)/2&{}\quad \text { if }{\widetilde{q}}\le q\le s, \\ \\ (a-q)q+(a-s+q)(s-q)&{}\quad \text { otherwise.} \end{array} \right. \end{aligned}$$
(57)

Assumption (35) requires that s belongs to the interval \(\left( a-c,\frac{2+\sqrt{3}}{2}\left( a-c\right) \right) \) and assumption \(\left( A4\right) \) requires that \(s<\mathbf {\ }\frac{a+c}{3}\). Under these assumptions, the fixed costs thresholds are ranked in the following order

$$\begin{aligned} F_{A}<F_{d}<{\overline{F}}. \end{aligned}$$
(58)

By solving the equation \(R\left( q,0\right) =R\left( q_{a},r\left( q_{a}\right) \right) \) for q, we obtain an explicit formula for \(\overline{ q}\):

$$\begin{aligned} {\overline{q}}=\frac{s}{2}-\frac{\sqrt{3}}{12}\sqrt{4s(2\left( a-c\right) -s)-\left( a-c\right) ^{2}}. \end{aligned}$$
(59)

One can easily check that \(4s(2\left( a-c\right) -s)-\left( a-c\right) ^{2}>0 \) for all \(s\in \left( a\!-\!c,\frac{2+\sqrt{3}}{2}\left( a\!-\!c\right) \right) \). Thus, \({\overline{q}}\) does exist for all \(s<\frac{2+\sqrt{3}}{2} \left( a-c\right) \), and we have \({\overline{q}}<q_{0}\). Further calculations show that \({\widetilde{q}}>{\overline{q}}\) holds only if \(F\ge F_{A}\).

Depending on the level of the recycler’s fixed costs, which determines the location of \({\widetilde{q}}\), there are three possibilities:

  1. (i)

    \({\widetilde{q}}<{\overline{q}}<q_{0}\). In the case where \(\widetilde{ q}<{\overline{q}}\), fixed costs are so small that \(F<F_{A}\). Then, \({\mathbf {V}} (q)\) has two local maxima at \({\widetilde{q}}\) and \(q_{a}\), with a global maximum at \(q_{a}\) because \({\widetilde{q}}<{\overline{q}}\) implies that \( R\left( {\widetilde{q}},0\right) <R\left( q_{a},r\left( q_{a}\right) \right) = {\mathbf {V}}(q_{a})\).

  2. (ii)

    \({\overline{q}}\le {\widetilde{q}}<q_{0}\). In this case, fixed costs are higher, so that \(F_{A}\le F<F_{d}\). As \({\overline{q}}\le {\widetilde{q}}\), we have \(R\left( q_{a},r\left( q_{a}\right) \right) \le \)\( R\left( {\widetilde{q}},0\right) \), which rules out \(q_{a}\) as a possible equilibrium. Moreover, \({\widetilde{q}}<q_{0}\) implies that \(F<F_{d}\), and one can check that \(F_{d}<{\overline{F}}\) when \(s<\frac{2+\sqrt{3}}{2}\left( a-c\right) \). Since \({\mathbf {V}}(q)\) is increasing on \(\left[ 0,{\widetilde{q}} \right] \), \({\mathbf {V}}(q)\) is maximized at \({\widetilde{q}}\).

  3. (iii)

    \(q_{0}\le {\widetilde{q}}\le s\). Now, fixed costs are even higher so that \(F_{d}\le F\le {\overline{F}}\). As \(q_{a}<q_{0}\), \({\mathbf {V}} (q)\) is decreasing on \(({\widetilde{q}},s]\), therefore \({\mathbf {V}}(q)\) achieves a unique maximum at \(q_{0}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ba, B.S., Mahenc, P. Is Recycling a Threat or an Opportunity for the Extractor of an Exhaustible Resource?. Environ Resource Econ 73, 1109–1134 (2019). https://doi.org/10.1007/s10640-018-0293-1

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10640-018-0293-1

Keywords

JEL Classification

Navigation