Skip to main content
Log in

Dispatching to parallel servers

Solutions of Poisson’s equation for first-policy improvement

  • Published:
Queueing Systems Aims and scope Submit manuscript

Abstract

Policy iteration techniques for multiple-server dispatching rely on the computation of value functions. In this context, we consider the continuous-space M/G/1-FCFS queue endowed with an arbitrarily designed cost function for the waiting times of the incoming jobs. The associated relative value function is a solution of Poisson’s equation for Markov chains, which in this work we solve in the Laplace transform domain by considering an ancillary, underlying stochastic process extended to (imaginary) negative backlog states. This construction enables us to issue closed-form relative value functions for polynomial and exponential cost functions and for piecewise compositions of the latter, in turn permitting the derivation of interval bounds for the relative value function in the form of power series or trigonometric sums. We review various cost approximation schemes and assess the convergence of the interval bounds these induce on the relative value function, namely Taylor expansions (divergent, except for a narrow class of entire functions with low orders of growth) and uniform approximation schemes (polynomials, trigonometric), which achieve optimal convergence rates over finite intervals. This study addresses all the steps to implementing dispatching policies for systems of parallel servers, from the specification of general cost functions toward the computation of interval bounds for the relative value functions and the exact implementation of the first-policy improvement step.

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
Fig. 5

Similar content being viewed by others

Notes

  1. The transition kernel \( P({u,x},\mathcal {S}) {\,=\,} Prob ((U_{n+1},X_{n+1})\,{\in }\, \mathcal {S}| (U_{n},X_{n}) \,{=}\, ( u,x)) \) satisfies

    $$\begin{aligned} P({u,x},\mathcal {U}\times \mathcal {X}) = \left\{ \begin{array}{llc} P({u+x},\{0\}) \, \mu _{X_{0}}(\mathcal {X}) &{} \text { if } &{} \mathcal {U}= \{0\}\\ P({u+x},\mathcal {U}) \, \mu _{X}(\mathcal {X}) &{} \text { if } &{} \mathcal {U}\not \supset \{0\} \end{array} \right\} , \end{aligned}$$
    (II1)

    where, for all \(u\ge 0\), one has \(P({u},[0,t]) = e^{-\lambda (u-t)}\) if \(t\in [0,u] \), and \( P({u},\mathbb {R}\setminus [0,u]) = 0\).

  2. Recall that the residue of a meromorphic function \(f\) at a pole \(a\) of order n is given by [10]

    $$\begin{aligned}\textstyle Res _{s=a}\left( {f({s})}\right) = \frac{1}{{(n-1)}!} \lim _{s\rightarrow a}\frac{d^{n-1}}{ds^{n-1}} \big [ (s-a)^n f({s}) \big ]. \end{aligned}$$
    (III6)
  3. Alternatively, notice that \( f_{a,n}=(-1)^{n} ({\delta ^n}/{\delta a^n})\, f_{a,0} \) if \(a\in {\mathcal {P}}_{W}\setminus \{0\}\). It follows from (CVF) and the Leibniz integral rule that, for \(a\in {\mathcal {P}}_{W}\setminus \{0\}\) and \(n>0\), \( c_{a:n}^{\prime }(u) = (-1)^{n} ({\delta ^n}/{\delta a^n}) c_{a:0}^{\prime }(u) = (-1)^{n} [{\lambda }/({1-\rho })] \, ({\delta ^n}/{\delta a^n}) [W^{*}(a) e^{-au} ] \), and the expressions for \(c_{a:n}^{\prime }\) can be derived by successive differentiations of \(c_{a:0}^{\prime }\). By continuity arguments, we also find, for \(n>0\), \( c_{0:n}^{\prime }(u) = (-1)^{n}\lim \nolimits _{a\rightarrow 0} ({\delta ^n}/{\delta a^n})[c_{a:0}^{\prime }(u)] \).

  4. Interval arithmetic. We use \([x]\equiv [x_1,x_2]\) to represent an interval on \(\mathbb {R}\). We write \([x]\in [\mathbb {R}]\) where \([\mathbb {R}]=\{[x_1,x_2]\, | \,x_1\le x_2;\, x_1,x_2\in \mathbb {R}\}\), \(a\in [x] \) iff \(x_1\le a \le x_2\), \(|[x]|=x_2-x_1\), and \(-[x] = [-x_2,-x_1] \). For \([x],[y]\in [\mathbb {R}]\) we have \([x]+[y] = [x_1+y_1,x_2+y_2] \), \([x]<[y]\) iff \(x_2<y_1\), and \([x]\le [y]\), \([x]>[y]\) and \([x]\ge [y]\) are defined similarly.

  5. Recall that the order of growth of an entire function \(f\) [30], defined by \(\varrho = \limsup _{r\rightarrow \infty }\ln \ln {\Vert f\Vert _{\infty ,r}}/\ln r\), where \(\Vert f\Vert _{\infty ,r} = \sup _{s}\{|f({s})|\, | \,|s|<r\}\), is the infimum of all m such that \(f({s})=O({{\,\mathrm{exp}\,}}(|s|^m))\), while the type of \(f\) is defined by \(\sigma = \limsup _{r\rightarrow \infty }\ln {\Vert f\Vert _{\infty ,r}}/ r^\varrho \). If \(\varrho =1\), then \(f\) is said to be of exponential type \(\sigma \).

  6. Piecewise continuous functions can be treated similarly by partitioning \(\mathbb {R}_{>0}\) into as many intervals as required by their discontinuities.

  7. In Algorithm 1, the first argument of \(\mathcal {A}_{i}\left( f;\cdot ,\cdot \right) \) (or \([\mathcal {A}_{i}]\left( [f];\cdot ,\cdot \right) \)) indicates the cost function \(f\) (resp. the interval function \([f]\)) for which the admission cost at server \(i\) is computed.

  8. In [28, §3], it is suggested to use the parameter values

    $$\begin{aligned} \textstyle \varrho _{n,0} {\,=\,} 1, \ \varrho _{n,1} {\,=\,} \cos (\frac{\pi }{n+2}), \ \varrho _{n,k} {\,=\,} \frac{\sum \nolimits _{q=0}^{n-k}\sin \left( \frac{q+1}{n+2}\pi \right) \sin \left( \frac{q+k+1}{n+2}\pi \right) }{ \sum \nolimits _{q=0}^{n} \sin ^2\left( \frac{q+1}{n+2}\pi \right) } for k=2,\dots ,n. \end{aligned}$$
    (IV6)

References

  1. Aalto, S., Virtamo, J.: Basic packet routing problem. In: 13th Nordic Teletrac Seminar, pp. 85–97 (1996)

  2. Arapostathis, A., Borkar, V., Fernández-Gaucherand, E., Ghosh, M., Marcus, S.: Discrete-time controlled Markov processes with average cost criterion: a survey. SIAM J. Control Optim. 31(2), 282–344 (1993). https://doi.org/10.1137/0331018

    Article  Google Scholar 

  3. Athreya, K.B., Ney, P.: A new approach to the limit theory of recurrent Markov chains. Trans. Am. Math. Soc. 245, 493–501 (1978)

    Article  Google Scholar 

  4. Bernstein, S.N.: Démonstration du Théoréme de Weierstrass fondée sur le calcul des Probabilités. Commun. Soc. Math. Kharkov 13(1), 1–2 (1912)

  5. Bertsekas, D.P.: Dynamic Programming and Optimal Control, vol. II, 3rd edn. Athena Scientific (2007)

  6. Bhulai, S.: On the value function of the M/Cox(r)/1 queue. J. Appl. Probab. 43, 363–376 (2006). https://doi.org/10.1239/jap/1152413728

    Article  Google Scholar 

  7. Bhulai, S., Spieksma, F.M.: On the uniqueness of solutions to the Poisson equations for average cost Markov chains with unbounded cost functions. Math. Methods Oper. Res. 58(2), 221–236 (2003)

    Article  Google Scholar 

  8. Bilenne, O.: Dispatching to parallel servers: solutions of Poisson’s equation for first-policy improvement—Python 2.7.16 code for reproducing the numerical experiments (2021). https://github.com/barafundle/poisson

  9. Bilenne, O.: Dispatching to parallel servers: solutions of Poisson’s equation for first-policy improvement—suplementary material (2021) https://urldefense.proofpoint.com/v2/url?u=https-3A__hal.archives-2Douvertes.fr_hal-2D02925284v3_file_QUESTA20-5FSupplementaryMaterialV2.pdf&d=DwIFaQ&c=vh6FgFnduejNhPPD0fl_yRaSfZy8CWbWnIf4XJhSqx8&r=cijxKIUfIjh6xB35XSxKelnSNfz2185wGO_qFr-DFH8&m=FfTwFytsZ4k9bNHBptyQhVzTrult0sM7a18bj_inDXU&s=xNunUEXId3Dl61488AII1dv910Vbft8wIJH82YtI9EE&e=

  10. Brown, J.: Complex Variables and Applications, 9th edn. McGraw-Hill Education, New York (2014)

    Google Scholar 

  11. Corless, R.M., Gonnet, G.H., Hare, D.E.G., Jeffrey, D.J., Knuth, D.E.: On the Lambert W function. Adv. Comput. Math. 5(1), 329–359 (1996). https://doi.org/10.1007/BF02124750

    Article  Google Scholar 

  12. De Turck, K., De Clercq, S., Wittevrongel, S., Bruneel, H., Fiems, D.: Transform-domain solutions of Poisson’s equation with applications to the asymptotic variance. In: Al-Begain, K., Fiems, D., Vincent, J.M. (eds.) Analytical and stochastic modeling techniques and applications, pp. 227–239. Springer, Berlin Heidelberg (2012)

  13. Gallager, R.G.: Stochastic Processes?: Theory for Applications. Cambridge University Press, Cambridge (2013)

    Book  Google Scholar 

  14. Glynn, P.W.: Poisson’s equation for the recurrent M/G/1 queue. Adv. Appl. Probab. 26(4), 1044–1062 (1994). https://doi.org/10.2307/1427904

  15. Glynn, P.W., Meyn, S.P.: A Liapounov bound for solutions of the Poisson equation. Ann. Probab. 24(2), 916–931 (1996). https://doi.org/10.1214/aop/1039639370

    Article  Google Scholar 

  16. Gross, D., Harris, C.M.: Fundamentals of Queueing Theory. Wiley, New York, Chichester, Weinheim (1998)

    Google Scholar 

  17. Howard, R.A.: Dynamic Programming and Markov Processes. MIT Press, Cambridge (1960)

    Google Scholar 

  18. Hyytiä, E., Aalto, S., Penttinen, A., Virtamo, J.: On the value function of the M/G/1 FCFS and LCFS queues. J. Appl. Probab. 49(4), 1052–1071 (2012)

    Article  Google Scholar 

  19. Hyytiä, E., Penttinen, A., Aalto, S.: Size- and state-aware dispatching problem with queue-specific job sizes. Eur. J. Oper. Res. 217(2), 357–370 (2012)

    Article  Google Scholar 

  20. Hyytiä, E., Righter, R., Aalto, S.: Task assignment in a heterogeneous server farm with switching delays and general energy-aware cost structure. Perform. Eval. 75–76, 17–35 (2014)

    Article  Google Scholar 

  21. Hyytiä, E., Righter, R., Bilenne, O., Wu, X.: Dispatching fixed-sized jobs with multiple deadlines to parallel heterogeneous servers. Perform. Eval. 114(Supplement C), 32–44 (2017). https://doi.org/10.1016/j.peva.2017.04.003

  22. Hyytiä, E., Righter, R., Virtamo, J., Viitasaari, L.: On value functions for FCFS queues with batch arrivals and general cost structures. Perform. Eval. 138, 102083 (2020). https://doi.org/10.1016/j.peva.2020.102083

    Article  Google Scholar 

  23. Hyytiä, E., Righter, R., Aalto, S.: Task assignment in a heterogeneous server farm with switching delays and general energy-aware cost structure. Perform. Eval. 75–76, 17–35 (2014). https://doi.org/10.1016/j.peva.2014.01.002

    Article  Google Scholar 

  24. Hyytiä, E., Virtamo, J., Aalto, S., Penttinen, A.: M/M/1-PS queue and size-aware task assignment. Perform. Eval. 68(11), 1136–1148 (2011). https://doi.org/10.1016/j.peva.2011.07.011. (Special Issue: Performance 2011)

    Article  Google Scholar 

  25. Sennott, I.L.: Average cost optimal stationary policies in infinite state Markov decision processes with unbounded costs. Oper. Res. 37, 626–633 (1989). https://doi.org/10.1287/opre.37.4.626

    Article  Google Scholar 

  26. Khintchine, A.Y.: Mathematical theory of a stationary queue. Mat. Sb. 39(4), 73–84 (1932)

    Google Scholar 

  27. Koralov, L., Sinai, Y.: Theory of Probability and Random Processes. Universitext. Springer, Berlin, Heidelberg (2007)

    Book  Google Scholar 

  28. Korovkin, P.: Linear Operators and Approximation Theory. Hindustan Publishing Corporation, International Monographs on Advanced Mathematics & Physics (1960)

    Google Scholar 

  29. Krishnan, K.R.: Joining the right queue: a Markov decision-rule. Proc. IEEE Confer. Decis. Control 26, 1863–1868 (1987). https://doi.org/10.1109/CDC.1987.272835

    Article  Google Scholar 

  30. Levin, B.: Lectures on Entire Functions. American Mathematical Society, Providence (1996)

    Book  Google Scholar 

  31. Meyn, S.P.: Convergence of the policy iteration algorithm with applications to queueing networks and their fluid models. Proc. IEEE Confer. Decis. Control 1, 366–371 (1996). https://doi.org/10.1109/CDC.1996.574337

    Article  Google Scholar 

  32. Meyn, S.P.: The policy iteration algorithm for average reward Markov decision processes with general state space. IEEE Trans. Autom. Control 42(12), 1663–1680 (1997). https://doi.org/10.1109/9.650016

    Article  Google Scholar 

  33. Mitrinović, D., Kečkić, J.: The Cauchy method of residues: theory and applications. Math. Appl. D. Reidel Publishing Company, Dordrecht, Holland (1984)

    Book  Google Scholar 

  34. Neveu, J.: Potentiel markovien récurrent des chaînes de Harris. Ann. Inst. Four. 22(2), 85–130 (1972). https://doi.org/10.5802/aif.414

    Article  Google Scholar 

  35. Nummelin, E.: On the poisson equation in the potential theory of a single kernel. Math. Scand. 68, 59–82 (1991). https://doi.org/10.7146/math.scand.a-12346

    Article  Google Scholar 

  36. Ott, T.J., Krishnan, K.R.: Separable routing: a scheme for state-dependent routing of circuit switched telephone traffic. Ann. Oper. Res. 35(1–4), 43–68 (1992). https://doi.org/10.1007/BF02023090

    Article  Google Scholar 

  37. Pollaczek, F.: Über eine Aufgabe der Wahrscheinlichkeitstheorie. I. Math. Z. 32(1), 64–100 (1930). https://doi.org/10.1007/BF01194620

    Article  Google Scholar 

  38. Rivlin, T.J.: An Introduction to the Approximation of Functions. Dover Publications, New York (1969)

    Google Scholar 

  39. Sassen, S., Tijms, H., Nobel, R.: A heuristic rule for routing customers to parallel servers. Stat. Neerl. 51, 107–121 (2001). https://doi.org/10.1111/1467-9574.00040

    Article  Google Scholar 

  40. Tijms, H.: A First Course in Stochastic Models. Wiley (2003)

  41. Welch, P.D.: On a generalized M/G/1 queuing process in which the first customer of each busy period receives exceptional service. Oper. Res. 12(5), 736–752 (1964). https://doi.org/10.1287/opre.12.5.736

    Article  Google Scholar 

  42. Wijngaard, J.: Decomposition for dynamic programming in production and inventory control. Eng. Process. Econ. 4(2), 385–388 (1979). https://doi.org/10.1016/0377-841X(79)90051-2

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Olivier Bilenne.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

The author acknowledges support from the French National Research Agency (Project ORACLESS, ANR-16-CE33-0004-01). Part of this work was completed at the Department of Communications and Networking, Aalto University, Espoo, Finland, with support from the Academy of Finland in the project FQ4BD (Grant No. 296206). The author is now with the Department of Data Science and Knowledge Engineering, Maastricht University, Netherlands (E-mail: o.bilenne@maastrichtuniversity.nl)

Appendices

On policy iteration and first-policy improvement

Recall the system depicted in Fig. 1, where jobs, arriving according to a Poisson process with rate \(\lambda \), are dispatched upon arrival toward one of the \(k\) servers (\(1,\dots ,k\)) selected by a (possibly random) dispatching policy \(\pi (u,x)\), where \(u=(u_{1},\dots ,u_{k})\in \mathbb {R}_{{\ge 0}}^k\) denotes the server backlog vector and \(x=(x_{1},\dots ,x_{k})\in \mathbb {R}_{{\ge 0}}^k\) are the prospective service times of an incoming job at the servers. By taking snapshots at initial time \(n=0\) and at the job arrival times (\(n=1,2,3,\dots \)), the continuous-time system reduces to a MDP, \((\Phi ^{\pi }_{n})_{n\in \mathbb {N}}\), with state \(\Phi ^{\pi }_{n} = ( U_{n} , X_{n} ) \in \Omega \equiv \mathbb {R}_{{\ge 0}}^k\times \mathbb {R}_{{\ge 0}}^k\), where \(X_{n}\) is the service time vector of the \(n\)th job and \(U_{n}\) is the backlog of the system at the time of arrival, and with transition probability kernel \(P=(P_{1},\dots ,P_{k})\) such that, for any \(n\) and every \((u,x)\in \Omega \), \(\mathcal {S}\subset \Omega \),

$$\begin{aligned} \begin{array}{c} P_{i}(u,x,\mathcal {S}) {:}{=}Prob (\Phi ^{\pi }_{n+1}\in \mathcal {S}| \Phi ^{\pi }_{n} = ( u,x),\, \pi (u,x) = i) = P_{i}((u_{1},\dots ,u_{i}+x,\dots ,u_{k}) , 0,\mathcal {S}) . \end{array} \end{aligned}$$

Assume that the performance of the system is measured by a cost function \(f\), where \(f(i,u,x) \equiv f_{i}(u_{i}) \, \mathbf {1}_{\mathbb {R}_{>0}}(x) \) models a penalty incurred when a job with service time \(x\) joins server \(i\), given backlog state \(u=(u_{1},\dots ,u_{k})\). We would like to minimize the long-run average cost, defined by

$$\begin{aligned} J_\pi = \limsup _{N\rightarrow \infty } \frac{1}{N} \sum _{n=1}^{N} \mathbb {E}\left[ f\left( \pi (\Phi ^{\pi }_{n}) ,\Phi ^{\pi }_{n} \right) \right] , \end{aligned}$$

independently of \(\Phi ^{\pi }_{0}\). The optimality equations of the system are

$$\begin{aligned} g({u,x})&\textstyle = \min _i\left[ f_{i}(u_{i}) + P_{i} g({u,x})\right] - \varsigma , \end{aligned}$$
(OEa)
$$\begin{aligned} \pi (u,x)&\textstyle \in \arg \min _i\left[ f_{i}(u_{i}) + P_{i} g({u,x})\right] , \end{aligned}$$
(OEb)

where \( P_{i} g({u,x}) = \int _\Omega g({t,y}) P_{i}(u,x,d(t,y)) \equiv P_{i} g({(u_{1},\dots ,u_{i}+x_{i},\dots ,u_{k}),0}) . \) If one can find \(\varsigma ^*>0\), a policy \(\pi ^*\), and an integrable function \(g\) such that \((g,\varsigma ^*)\) solves (OEa) and \(\pi ^*\) satisfies (OEb), then \(\pi ^*\) is the optimal policy and

is the optimal cost of the system [2, 25, 31, 32]. The policy iteration algorithm for solving (OE) can be described as follows [17]: Given an initial policy \(\pi ^{(0)}\), find, for \(k\ge 0\), a function \({g}^{(k)}\), a mean cost \(\varsigma ^{(k)}\), and a policy \(\pi ^{(k+1)}\) satisfying

$$\begin{aligned} \textstyle {g}^{(k)}(u,x)&\textstyle = f( \pi ^{(k)}{(u,x)} ,u,x) + P_{\pi ^{(k)}{(u,x)}} {g}^{(k)}(u,x) - \varsigma ^{(k)}, \end{aligned}$$
(PIa)
$$\begin{aligned} \textstyle \pi ^{(k+1)}{(u,x)}&\textstyle \in \arg \min _i\left[ { f_{i}(u_{i}) } + {v}^{(k)}(u_{1},\dots ,u_{i}+x_{i},\dots ,u_{k}) \right] , \end{aligned}$$
(PIb)

where (PIa) is the policy evaluation step, (PIb) is the policy improvement step and \( {v}^{(k)}(u) := \int _\Omega {g}^{(k)}(t,y) \, P({u,0},d(t,y)) - \varsigma ^{(k)} \) defines the relative value function under policy \(\pi ^{(k)}\). Under favorable conditions, \(( \pi ^{(k)}, \varsigma ^{(k)})\) eventually converges toward a solution \( (\pi ^*,\varsigma ^*)\) [31, 32]. Solving (PIb), however, is generally difficult.

The first iteration of (PI) may still be implemented easily if the initial policy \(\pi ^{(0)}\equiv \pi \) is a random Bernoulli split between the servers. In that case, the Poisson process separates into \(k\) independent Poisson processes with rates \(\lambda _{1},\dots ,\lambda _{k}\), where \(\lambda _{1}+\cdots +\lambda _{k}=\lambda \), and the multiple-server system decomposes into \(k\) independent M/G/1 queues with arrival rates \(\lambda _{1},\dots ,\lambda _{k}\) and transition probability kernels \(P^{{(\lambda _{1})}},\ldots ,P^{{(\lambda _{k})}}\). The relative value function under the initial random policy then takes the additively separable form \(v(u) = \sum _{i=1}^{k} v_{i}(u_{i})\), where \( v_{i}(u_{i}) {:}{=}P^{{(\lambda _{i})}} g_{i}({u_{i},0}) - \varsigma _{i} \) is the relative value function at server \(i\), and \(g_{i}\) satisfies

$$\begin{aligned} g_{i}({u_{i},x_{i}}) = P^{{(\lambda _{i})}} g_{i}({u_{i},x_{i}}) + f_{i}(u_{i}) - \varsigma _{i} \qquad (i=1,\dots ,k). \end{aligned}$$
(FPIa)

Since \( f_{i}(u_{i}) + v(u_{1},\dots ,u_{i}+x_{i},\dots ,u_{k}) = f_{i}(u_{i}) + v(u) + v_{i}(u_{i}+x_{i}) - v_{i}(u_{i}) \), (PIb) reduces to

$$\begin{aligned} \hat{\pi }{(u,x)} \in \arg \min _i\mathcal {A}_{i}(u,x) , \end{aligned}$$
(FPIb)

where \( \mathcal {A}_{i}(u,x) = f_{i}(u_{i}) + v_{i}(u_{i}+x_{i}) - v_{i}(u_{i}) \) is the admission cost at server \(i\). The first-policy improvement approach then consists in stopping the PI algorithm after a single iteration, by solving (FPIa)–(FPIb). Observe that (FPIa) is an instance of the Poisson equation \(g= Pg+ f\) under \(\int f(u)\, \nu ({du}) = 0\), where \(\nu \) denotes the nontrivial measure invariant for the transition kernel P (i.e., \(P \nu = \nu \)) [3, 34, 35]. In (FPIa), \(\nu \) coincides with the asymptotic probability measure of the waiting times at server i.

Characterization of the value function

Before showing Propositions 12, and Theorem 1, we characterize \(W^{*}\) in the complex plane.

Proposition B.1

(Analyticity of \(W^{*}\) and pole location) Under Assumption 1:

  1. (i)

    The dominant singularity \(p_{W}\) of \(W^{*}\) (i.e., that with largest real value) is a pole with degree 1 lying on the negative real axis \(\mathbb {R}_{<0}\). The dominant singularity \(p_{X}\) of \(X^{*}\) is real, negative (possibly infinite) and satisfies \(p_{X}<p_{W}\). \(X^{*}\) is analytic on \(\{s\in \mathbb {C}\, | \,\mathfrak {R}\left( s\right) > p_{W}\}\).

  2. (ii)

    \(W^{*}\) is analytic on \(\{s\in \mathbb {C}_{0}\, | \,\mathfrak {R}\left( s\right) > p_{W}\}\), where \(\lim \nolimits _{s\rightarrow \infty }|W^{*}(s)|\le 1\).

  3. (iii)

    One can find \(\epsilon >0\) such that \(W^{*}\) is analytic on \(\{s\in \mathbb {C}_{0}\setminus \{p_{W}\}\, | \,\mathfrak {R}(s)>p_{W}-\epsilon \}\).

  4. (iv)

    \(W^{*}\) is analytic in a neighborhood of 0, where it can be rewritten as the series

    $$\begin{aligned} W^{*}(s) = \sum \nolimits _{k=0}^\infty w_{k}\, (-s)^k , \quad \forall s\in \{\sigma \in \mathbb {C}_{0}:|\sigma -a| < |p_{W}| \}, \end{aligned}$$
    (B1)

    in which the coefficients \(\{w_{k}\}\) are given by (III3) in Table 1, and satisfy \(w_{k}=\mathbb {E}\big [W^k\big ]/ {k}!\), for \(k\in \mathbb {N}\). The series \(\{w_{k}\}\) is asymptotically geometric with asymptotic rate \(|p_{W}|^{-1}\).

  5. (v)

    At any point \(a\in \mathbb {C}_{0}\) where \(W^{*}\) is analytic, \(W^{*}\) can be rewritten as the series

    $$\begin{aligned} W^{*}(s) = \sum \nolimits _{k=0}^{\infty } w_{{a}:{k}} (a-s)^{k} , \quad \forall s\in \{\sigma \in \mathbb {C}:|\sigma -a| < r_{a} \} , \end{aligned}$$
    (B2)

    where \(r_{a}\) denotes the distance from \(a\) to the closest singularity of \(W^{*}\). The coefficients \(\{w_{{a}:{k}}\}\) are given by (III4) in Table 1.

Next, we derive the identities of Sect. 3.1 for the value function.

Proof

(Proposition 1) Start the queue at state \(u\). The quantity \( V(u,t) \) appearing in (VF) can be rewritten, for any \(T\ge 0\) and for \(t\) large enough, as \( V(u,t) = V(u,T) + V(U({T}),t-T) \), where \(U({T})\) denotes the backlog observed after time \(T\). It follows from the Markov property of the system and from Definition (VF) of the value function that

$$\begin{aligned} v(u) = \mathbb {E}\big [ V(u,T) - \lambda \bar{f}T\big ] + \mathbb {E}\big [ v(U({T})) \big ] . \end{aligned}$$
(B3)

Now, consider the function

$$\begin{aligned} \textstyle g({u,x}) = \lim \nolimits _{N\rightarrow \infty }\{ \sum _{n=1}^{N} \mathbb {E}\left[ f(U_{n} ) -\bar{f}\right] \, | \,(U_{1},X_{1})=(u,x) \} + \bar{f}, \quad \forall u,x\in \mathbb {R}_{{\ge 0}},\nonumber \\ \end{aligned}$$
(B4)

which can be verified to satisfy Eq. (PE) by application of the Markov property to the MDP. The function \(g\) defined by (B4) can be seen as a discrete-time counterpart of the value function (VF), which follows from (B4) by using the convention \((U_{0},X_{0})=(u,0)\) and setting \(v(u)\equiv Pg({u,0}) - \bar{f}\) or, equivalently, from (B3) by defining \(T\) as the arrival time of the first job so that \(\mathbb {E}\big [ V(u,T) - \lambda \bar{f}T\big ] = -\bar{f}\) and \(\mathbb {E}\big [ v(U({T})) \big ] \equiv Pg({u,0})\). \(\square \)

Proof

(Theorem 1) A simple calculation reveals that \(\Updelta (u)=0\) if \(u<0\), and \( \textstyle \Updelta (u) = (P- \widehat{P}) g({u,0}) = \kappa e^{-\lambda u} \), for \(u\ge 0\), with \(\kappa \) satisfying

$$\begin{aligned} \textstyle \kappa = \mathbb {E}\big [ g({0,X_{0}}) - \lambda \int _{-\infty }^{0} g({u,X}) \, e^{\lambda u} \, du\big ] . \end{aligned}$$
(B5)

We characterize the extended value function \( \widehat{v}:u\in \mathbb {R}\mapsto \widehat{v}(u)=g({u,0}) - f(u) \) associated with some \(g\) solution of (PE’). Note that, by construction, \(\widehat{v}\) coincides with the value function on \(\mathbb {R}_{{\ge 0}}\), i.e., \(\widehat{v}(u) \equiv v(u)\) if \(u\ge 0\). Once \(\widehat{v}\) is known, it will be possible to recover \(g\) using

$$\begin{aligned} g({u,x}) = g({u+x,0}) - f(u+x) + f(u) = \widehat{v}(u+x) + f(u) . \end{aligned}$$
(B6)

Consider \(s\) in the region of absolute convergence of \(\mathcal {B}_{\widehat{v}}\), where the orders of integration in our developments may be permuted. The two-sided Laplace transform of \(\widehat{v}\) is given by

Solving the above equation for \(\mathcal {B}_{\widehat{v}}(s)\) and using \( W^{*}(-s) = 1+{\lambda \mathbb {E}\big [X^2\big ]}/[{2(1-\rho )}]s+o(s)\) yields, after computations,

(B7)

where \(h(s)\) has no singularities on \(\{s\in \mathbb {C}\, | \,\mathfrak {R}\left( s\right) <- p_{W}\}\), and \(\varepsilon \) is given by

$$\begin{aligned} \varepsilon = {\kappa }/{(1-\rho )}-{\hat{X}^{*}_{2}(\lambda )\bar{f}}/{(1-\rho )^2} , \end{aligned}$$
(B8)

with \( \hat{X}^{*}_{2}(\lambda ){:}{=}1-\lambda \mathbb {E}\big [X\big ]+\lambda ^2\mathbb {E}\big [X^2\big ]/2 \). Since \(\widehat{v}\) is expected to be asymptotically flat for \(u\rightarrow -\infty \), the \(- {\lambda \bar{f}}/[{(1-\rho )s^2}] \) term in (B7) is necessarily due to a term \( -{\lambda \bar{f}}/({1-\rho })\, u\) on \(\mathbb {R}_{{\ge 0}}\) in the backlog domain. By inverse transformation of (B7), we find

figure f

where \(r\) satisfies \(\mathcal {L}_{r(-\cdot )}(s) = -[{h(-s)+\varepsilon }]/{s}\). The general form for \(g\) follows from (B6), (\(\widehat{\text {S}}\)) and \(\widehat{v}(u) \equiv v(u)\) on \(\mathbb {R}_{{\ge 0}}\). The nonempty ROC of \(\mathcal {B}_{\partial _{+}\widehat{v}}\) follows from \(-p_{W}\in ROC (\mathcal {L}_{f})\).

It remains to show that the function \(r\) is identical for all solutions or, equivalently, that the quantity \(\kappa \) in (B5) is the same for all \(g\).

To see this, consider a solution \(g_{1}\) of (PE’) with associated value function \(\widehat{v}_{1}\) and jump \(\varepsilon _{1}\) at \(u=0\). The value function for every other solution \(g_{2}\) can be rewritten as \(\widehat{v}_{2}=\alpha + (\widehat{v}_{1} - \varepsilon _{2} + \varepsilon _{1} ) \mathbf {1}_{\mathbb {R}_{<0}}+ \widehat{v}_{1} \mathbf {1}_{\mathbb {R}_{{\ge 0}}}\), where \(\varepsilon _{1}\), \(\varepsilon _{2}\) and \(\alpha \) are constants. We show that \(\varepsilon _{1}=\varepsilon _{2}\). If we successively compute expression (B8) for \(\varepsilon _{1}\) and \(\varepsilon _{2}\), using (B5), (B6) and the extension of (PE), we get, after simplifications, \( \varepsilon _{2} = \varepsilon _{1} - ( \varepsilon _{1} - \varepsilon _{2} ) X^{*}(\lambda ) /{(1-\rho )} \). Exploiting twice the strict convexity of \(e^{-x}\), we find \( X^{*}(\lambda ) = \mathbb {E}\big [ e^{-\lambda X} \big ]> e^{-\lambda \mathbb {E}\big [X\big ] } > 1 - \lambda \mathbb {E}\big [X\big ] = 1-\rho \). Hence, \(X^{*}(\lambda ) /{(1-\rho )} \ne 1\) and, consequently, \(\varepsilon _{1}=\varepsilon _{2}\). \(\square \)

Proof

(Proposition 2) (i) Consider \(s\) in the region of absolute convergence of \(\mathcal {B}_{\widehat{v}}\). Since \(\mathcal {B}_{\partial _{+}{c}}(s)=s\mathcal {B}_{{c}}(s)\), (C) can be rewritten as

$$\begin{aligned}{}[s+ \lambda (1- X^{*}(-s)) ] \, \mathcal {B}_{c}(s) \overset{\text {\tiny {(PK)}}}{=} \lambda \mathcal {L}_{f}(s) + s\mathcal {L}_{r(-\cdot )}(-s) \left[ 1+ \lambda /s(1- X^{*}(-s)) \right] , \end{aligned}$$

while transformation of (\(\widehat{\text {S}}\)) gives \( \mathcal {B}_{c}(s)= \mathcal {B}_{\widehat{v}}(s) + {\lambda \bar{f}}/[{(1-\rho )s^2}] \). Besides,

$$\begin{aligned} X^{*}(-s) \mathcal {B}_{\widehat{v}}(s)= & {} \mathbb {E}\big [{sX}\big ]\int _{-\infty }^{+\infty } \widehat{v}(u) e^{-su} du= \mathbb {E}\big [ \int _{-\infty }^{+\infty } \widehat{v}(u) e^{-s(u-X)} du\big ]\\= & {} \mathbb {E}\big [ \int _{-\infty }^{+\infty } \widehat{v}(t+X) e^{-st} dt\big ] = \int _{-\infty }^{+\infty } \mathbb {E}\big [ \widehat{v}(t+X) \big ] e^{-st} dt= \mathcal {B}_{\mathbb {E}\big [\widehat{v}(\cdot +X)\big ]}(s). \end{aligned}$$

Combining the above with \(\mathcal {B}_{\partial _{+}\widehat{v}}(s)=s\mathcal {B}_{\widehat{v}}(s)\), we get, after computations,

$$\begin{aligned} \mathcal {B}_{\partial _{+}\widehat{v}}(s) = \lambda \left[ \mathcal {L}_{f}(s) -\bar{f}/s+ \mathcal {B}_{\mathbb {E}\big [\widehat{v}(\cdot +X)\big ]}(s) - \mathcal {B}_{\widehat{v}}(s)\right] + \tilde{h}(s) , \end{aligned}$$
(B9)

where \(\tilde{h}(s)\) shows no singularity on \(\{s\in \mathbb {C}\, | \,\mathfrak {R}\left( s\right) <- p_{W}\}\), and we have used Proposition B.1(i) and \( 1+ \lambda /s(1- X^{*}(-s)) = 1-\rho +o(1) \). Inverse Laplace transformation of (B9) then gives, at every \(u\ge 0\) where \(\widehat{v}\) is differentiable,

$$\begin{aligned} \widehat{v}^{\prime }(u) = \lambda \left( f(u) - \bar{f}+ \mathbb {E}\left[ \widehat{v}(u+X)- \widehat{v}(u) \right] \right) , \end{aligned}$$
(B10)

which holds for almost every \(u>0\) by piecewise continuity of \(f\). Since by construction \(v(u)\equiv \widehat{v}(u)\) for \(u\ge 0\), we find (DE). From Theorem 1, we have

$$\begin{aligned} \textstyle \widehat{v}(0) = g({0,0}) - f(0) \overset{\text {\tiny {(PE')}}}{=} Pg({0,0}) - \bar{f}\overset{\text {\tiny {(II1)}}}{=} \mathbb {E}\big [\widehat{v}(X_{0})\big ] + f(0) - \bar{f}, \end{aligned}$$
(B11)

which yields (BCa). Finally, we find (BCb) by taking the limit of (B10) as \(u\rightarrow 0^+\),

(ii) Equation (S) follows directly from (\(\widehat{\text {S}}\)) and the fact that \(v(u)\equiv \widehat{v}(u)\) for \(u\ge 0\). It remains to compute \(\partial _{+}c\). From Theorem 1, we get

(B12)

where we have used \(f(u)=0\) if \(u<0\). Equation (C) follows by inversion of (B12). \(\square \)

Splitting schemes for the Pollaczek–Khinchin formula

This section includes further details on the transform splitting scheme (III12), which proves very useful when computing value functions with discrete service time distributions. See also [9] for a step-by-step derivation of the core function in the case of jobs with identical service times.

1.1 Discrete service time distributions

Consider the M/D/1 queue, where all the jobs have equal service time \(x\). In this scenario, the transform \(W^{*}\) is given by (PK) with \(X^{*}(s)=e^{-sx}\) and can be rewritten as

$$\begin{aligned} \begin{array}{ll} W^{*}(s) = [\Upsilon (-s)]^m W^{*}(s) +\frac{(1-\lambda x)s}{s- \lambda } \sum \nolimits _{k=0}^{m-1} [\Upsilon (-s)]^k ,&\quad \forall m\in \mathbb {N}_{>0}, \end{array} \end{aligned}$$
(C1)

where \(\Upsilon (s) {:}{=}[{\lambda }/({s+\lambda })]\,e^{sx}\). It can be seen that (III12) holds if

$$\begin{aligned} \textstyle W^{l}_{u}(s) = [\Upsilon (-s)]^ {\tilde{m}({u})} W^{*}(s) ,\quad W^{r}_{u}(s) = \frac{s(1-\rho ) }{s-\lambda } \sum \nolimits _{k=0}^{\tilde{m}({u})-1} [\Upsilon (-s)]^k , \end{aligned}$$

with \( \tilde{m}({u})=\left\lceil ({\tau -u})/{x}\right\rceil \).

1.2 Degenerate cases

The decomposition scheme (C1) is not possible for all discrete service time distributions. Consider, for instance, the geometric service time distribution \( F_{X}({u})= (e^{\varsigma }-1) \sum \nolimits _{k=1}^{\infty } e^{- k\varsigma } \theta (u-kx) \) for \( u\in \mathbb {R}_{{\ge 0}}\), where \(x>0\) and \(\lambda <(1-e^{-\varsigma })/x\). We have \(\mathbb {E}\big [ X\big ] = x/(1-e^{-\varsigma })\), \(X^{*}(s) = (e^{\varsigma }-1)/(e^{\varsigma +sx}-1) \), and \(W^{*}\) degenerates into

$$\begin{aligned} W^{*}(s) = \frac{ b(e^{\varsigma +sx}-1) s}{(s- \lambda ) e^{\varsigma +sx} -s+\lambda e^{\varsigma } } = b-s(e^{\varsigma +sx}-1) \, f(s) , \end{aligned}$$
(C2)

where \(b= ({1-e^{-\varsigma }-\lambda x})/({1-e^{-\varsigma }})\) and \(f(s) = [(\lambda -s) e^{\varsigma +sx} +s-\lambda e^{\varsigma } ]^{-1} \). Although \(f(s)\) decreases like \(O(r^{-1})\) as \(|s|\rightarrow \infty \) (i.e., not fast enough for counterclockwise integration along \(\mathcal {C}_{r}\)), it decomposes as follows:

$$\begin{aligned} \textstyle f(s) = [\tilde{\Upsilon }(-s)]^{m } f(s) + \sum \nolimits _{k=1}^{m} [\tilde{\Upsilon }(-s)]^{k} /({\lambda e^{\varsigma }-s}), \quad (m=1,2,\dots ), \end{aligned}$$
(C3)

where \( \textstyle \tilde{\Upsilon }(s)= ({ s+\lambda e^{\varsigma }})/{(s+ \lambda ) } \, e^{sx-\varsigma } \) is \(O( e^{\mathfrak {R}(s)x} )\) with just one pole at \(-\lambda \). By distributing (C2) and using (C3) twice with parameters \(m+1\) and m, we find, after computations, that (III12) holds if we set

$$\begin{aligned} \begin{array}{rcl} W^{l}_{u}(s) &{}=&{} \lambda b\frac{s( e^{\varsigma }-1 )( \lambda e^{\varsigma }-s)^{\tilde{m}_{l}({u})-1}}{(\lambda -s)^{\tilde{m}({u})}\, [ s- \lambda -(s-\lambda e^{\varsigma }) e^{-(sx+\varsigma )} ] } \, e^{-\tilde{m}({u})(sx+\varsigma )},\\ W^{r}_{u}(s) &{}=&{} \frac{s-b}{s-\lambda } + \lambda b\frac{ (e^{\varsigma }-1 ) -s}{(s-\lambda ) (s-\lambda e^{\varsigma })} \sum \nolimits _{k=1}^{\tilde{m}({u})-1} \big (\frac{s-\lambda e^{\varsigma }}{s-\lambda } \big )^{k} e^{-k(sx+\varsigma )} , \end{array} \end{aligned}$$

with \( \tilde{m}({u})=\lceil ({\tau -u})/{x}\rceil \).

System interpretation of Taylor series CVF

In view of (C), the core function can be seen as the answer to the input \(f\) of a linear filter with transfer function \({\lambda }/{(1-\rho )} \,W^{*}(-s)\). Similarly, the Taylor series (IV3) for the CVF of an analytic cost function is reminiscent of digital filtering in signal processing. Indeed, for \(n\in \mathbb {N}\), consider the vector of derivatives \(\tilde{f}^{(n)}=(\tilde{f}_{0},\dots , \tilde{f}_{n-1} )\), where \(\tilde{f}_{k} ={f^{(k)}(0)}\), and the vector \(\tilde{c}^{(n)}=(\tilde{c}^{(n)}_{0}, \dots ,\tilde{c}^{(n)}_{n-1})\), where \(\tilde{c}^{(n)}_{k}\) is defined by (IV2). Using the matrix inversion lemma, one shows that the Toeplitz, upper triangular matrices

satisfy \( W^{(n)}({I^{(n)}}/{\lambda }-X^{(n)}) = I^{(n)} \) for all \(n\), where \(I^{(n)}\) denotes the identity matrix. Besides, (IV2) rewrites in matrix form as \( \tilde{c}^{(n)}=W^{(n)} \tilde{f}^{(n)} \).

As \(n\rightarrow \infty \), \(\{\tilde{c}_{t}\}\) becomes the output (at nonnegative times) of the cross-correlation of \(\{\tilde{f}_{t}\}\) with the sequence defined by \(\tilde{h}[t]= {\lambda }/({1-\rho })\,w_{t} \) for \(t\ge 0\), thus giving us an interpretation for analytic functions of Proposition 2(ii), where \(c^{\prime }\) was obtained by cross-correlation of \(f(u)\) with \(\lambda /(1-\rho ) F^{\prime }_{W}({u})\). Similarly, (IV3b) expresses \(c^{\prime }(u)\) as the cross-correlation of the sequence of derivatives of \(f\) at \(u\) with the sequence \(\tilde{h}[t]\).

From our observation it follows that the Z-transforms of the sequences satisfy

$$\begin{aligned} \mathcal {Z}_{\tilde{c}^{(\infty )}}(z) = \mathcal {Z}_{\tilde{h}}({1}/{z})\, \mathcal {Z}_{\tilde{f}^{(\infty )}}(z) , \end{aligned}$$
(D1)

where \( \mathcal {Z}_{h}(z) = \sum \nolimits _{k=0}^{\infty } h[k] z^{-k}\) denotes the Z-transform of a sequence \(h[t]\). The vector \(\tilde{c}^{(\infty )}\) can be recovered from (D1) by inverse Z-transform, provided that the regions of convergence of \(\mathcal {Z}_{\tilde{f}^{(\infty )}}(z)\) and \(\mathcal {Z}_{\tilde{h}}({1}/{z})\) intersect on a nonempty circular band – this condition is to be linked to those of Theorem 2(i).

Conversely, inverting \(W^{(n)}\) yields \(\tilde{f}^{(n)}=({I^{(n)}}/{\lambda }-X^{(n)}) \tilde{c}^{(n)}\) and, as \(n\rightarrow \infty \),

$$\begin{aligned} \tilde{f}_{k} = {\tilde{c}_{k}}/{\lambda } - \sum \nolimits _{t=0}^\infty [{\tilde{c}_{t+k}}/{{(t+1)}!}]\, \mathbb {E}\big [X^{t+1}\big ] , \end{aligned}$$
(D2)

which provides us with a converse for Theorem 2, where the source cost function of a given core function with germ \(\{\tilde{c}_{k}\}\) can be recovered from \(c\) through (D2), on the condition that \(c\) grows slower than the exponential type \(|p_{X}|\) – where \(p_{X}\) denotes the dominant pole of \(X\) – , in which case \(({I^{(n)}}/{\lambda }-X^{(n)})\,\tilde{c}^{(n)}\) converges as \(n\rightarrow \infty \). Similarly, \( \mathcal {Z}_{\tilde{f}^{(\infty )}}(z) = (\mathcal {Z}_{\tilde{h}}({1}/{z}))^{-1} \mathcal {Z}_{\tilde{c}^{(\infty )}}(z) \), where \((\mathcal {Z}_{\tilde{h}}(z))^{-1}\) is the Z-transform of \(\delta [{t}]/\lambda - \mathbb {E}\big [X^{t+1}\big ]/{(t+1)}! \).

Proofs and auxiliary results

Proof

(Theorem 2) (i) If \(\varrho \) is the order of growth of the entire cost function \(f\), and \(\sigma \) is its type, then for any \(\epsilon >0\) there is \(k_\epsilon <\infty \) such that [30, Lecture 1]

$$\begin{aligned} \begin{array}{ll} \frac{1}{{{k}}!}|f^{({k})}(0)| < \big (\frac{e (\varrho +\epsilon )}{{k}}\big )^{\frac{{k}}{\varrho +\epsilon }} ,&\quad \forall {k}>k_\epsilon , \end{array} \end{aligned}$$
(E1a)
$$\begin{aligned} \begin{array}{ll} \frac{1}{{{k}}!}|f^{({k})}(0)| < \big (\frac{e (\sigma +\epsilon ) \varrho }{{k}}\big )^{\frac{{k}}{\varrho }} ,&\quad \forall {k}>k_\epsilon . \end{array} \end{aligned}$$
(E1b)

Consider the quantity \(\tilde{c}_{{k}}= \sum \nolimits _{q=0}^{\infty } w_{q} \, f^{(k+q)}(0)\) introduced in (IV2), as well as

$$\begin{aligned} \begin{array}{ll} \bar{c}_{{k}}=\frac{\lambda }{1-\rho } \sum \nolimits _{q=0}^{\infty } w_{q} \left| f^{({k}+q)}(0)\right| ,&\quad \forall {k}\in \mathbb {N}. \end{array} \end{aligned}$$
(E2)

Recall from Proposition B.1 (iv) in Appendix B that \(\lim \nolimits _{k\rightarrow \infty } w_{k+1}/w_{k} = |p_{W}|^{-1}\). Besides, it can be seen (for example, using Stirling’s approximation for the factorial) that

$$\begin{aligned} \begin{array}{ll} \lim \nolimits _{{k}\rightarrow \infty } \frac{ {({k}+1)}! \big (\frac{e s r}{{k}+{l}+1}\big )^{\frac{{k}+{l}+1}{r}} }{ {{k}}! \big (\frac{e s r}{{k}+{l}}\big )^{\frac{{k}+{l}}{r}} } = \Bigg \lbrace \begin{array}{cl} 0,&{}if r<1 \\ s,&{}if r=1 \\ \infty ,&{}if r>1 \end{array}\Bigg \rbrace ,&\quad \forall {l}\in \mathbb {N}. \end{array} \end{aligned}$$
(E3)

Equations (E1a) and (E1b) tell us that, under the assumptions of (i) and by taking \(\epsilon \) sufficiently small, one can find a dominant series for \(\tilde{c}_{{k}}\) and \(\bar{c}_{{k}}\) that successfully passes the ratio test for convergence due to (E3), so that both \(\tilde{c}_{{k}}\) and \(\bar{c}_{{k}}\) are finite for all k. The finiteness of \(\bar{c}_{{k}}\) allows us to interchange the integration order in the computation of \(\tilde{c}_{{k}}\). Noting that \(w_{q}=\mathbb {E}\big [W^q\big ]/ {q}!\) for all q [cf. Proposition B.1(iv)], we apply Fubini’s theorem and find, for \({k}\in \mathbb {N}_{\ge 0}\),

(E4)

Similarly, we introduce, for \({k}\in \mathbb {N}\),

(E5)

and \(\hat{c}_{{k}}\) is finite as well. Suppose now that \(|({d^{{k}}}/ {du^{{k}}})f(0)| < {{{k}}!} ({e s r}/{{k}})^{{{k}}/{r}}\) for \({k}>k_\epsilon \)–in the case (i), this holds either for some \(r<1\) or for \(r=1\) and some finite s–, and consider the sequence

$$\begin{aligned} \textstyle \beta _{{k}}= \frac{\lambda }{1-\rho } \sum \nolimits _{q=0}^{\infty } {(q+{k})}! \, w_{q} \big (\frac{e s r}{q+{k}}\big )^{\frac{q+{k}}{r}} , \quad \forall {k}\in \mathbb {N}. \end{aligned}$$
(E6)

It is easy to see that the three sequences \(\sum \nolimits _{{k}=0}^{\infty }\tilde{c}_{{k}} { u^{{k}+1}}/{{({k}+1)}!} \), \(\sum \nolimits _{{k}=0}^{\infty }\bar{c}_{{k}} { u^{{k}+1}}/{{({k}+1)}!} \) and \(\sum \nolimits _{{k}=0}^{\infty }\hat{c}_{{k}} { u^{{k}+1}}/{{({k}+1)}!} \) converge wherever \(\sum \nolimits _{{k}=0}^{\infty } \beta _{{k}} { u^{{k}+1}}/{{({k}+1)}!} \) is convergent. Besides,

(E7)

Under the conditions of (i), we infer from E3 that the expression between brackets in (E7) tends to a finite quantity not larger than s, so that for any \(\nu >0\) one can find a \(k_\nu \) such that \(\beta _{{k}+1}\le (\beta _{k_\nu +1}-\beta _{k_\nu }) + (s+\nu ) \beta _{{k}} \) for \({k}>k_\nu \). It follows from the ratio test that \(\sum \nolimits _{{k}=0}^{\infty } \beta _{{k}} { \xi ^{{k}}}/{{{k}}!} \) converges for \(\xi \in \mathbb {R}_{{\ge 0}}\), and so do \(\sum \nolimits _{{k}=0}^{\infty }\tilde{c}_{{k}} { \xi ^{{k}}}/{{{k}}!} = \psi (u)\), \(\sum \nolimits _{{k}=0}^{\infty }\bar{c}_{{k}} {\xi ^{{k}}}/{{{k}}!} \) and \(\sum \nolimits _{{k}=0}^{\infty }\hat{c}_{{k}} {\xi ^{{k}}}/{{{k}}!} \). This last conclusion, together with (E4), (E5), and Fubini’s theorem applied to set of natural numbers with the counting measure, yields, for \(u\in \mathbb {R}_{{\ge 0}}\),

where the last result follows from Proposition 2(ii). Since

$$\begin{aligned} \sum \nolimits _{{k}=0}^{\infty }\bar{c}_{{k}} \frac{\xi ^{{k}}}{{{k}}!} = \frac{\lambda }{1-\rho } \sum \nolimits _{{k}=0}^{\infty } \left( \sum \nolimits _{q=0}^{\infty }| w_{q} \, f^{({k}+q)}(0) \frac{ \xi ^{{k}}}{{{k}}!}|\right) <\infty , \quad \forall \xi \in \mathbb {R}_{{\ge 0}}, \end{aligned}$$

Fubini’s theorem applies and one may interchange the order of summation in (IV3a):

which holds for \(u\in \mathbb {R}_{{\ge 0}}\).

(ii) Similarly, for any \(\epsilon >0\), one can find growing sequences of naturals \(\{l_{{k}}\}\) and \(\{m_{{k}}\}\) such that [30, Lecture 1]

$$\begin{aligned} \begin{array}{ll} \frac{|f^{(l_{{k}})}(0)| }{{l_{{k}}}!} > \big (\frac{e (\varrho -\epsilon )}{l_{{k}}}\big )^{\frac{l_{{k}}}{\varrho -\epsilon }} ,&\quad ({k}\in \mathbb {N}_{\ge 0}), \end{array} \end{aligned}$$
(E8a)
$$\begin{aligned} \begin{array}{ll} \frac{|f^{(m_{{k}})}(0)| }{{m_{{k}}}!} > \big (\frac{e (\sigma -\epsilon ) \varrho }{m_{{k}}}\big )^{\frac{m_{{k}}}{\varrho }} ,&\quad ({k}\in \mathbb {N}_{\ge 0}). \end{array} \end{aligned}$$
(E8b)

Recall the series \(\tilde{c}_{{k}}\) defined in (IV2). By taking \(\epsilon \) sufficiently small in (E8a) and (E8b) and using (E3), we find that the asymptotic ratio between the moduli of two terms of (IV2) with respective indices \(l_{q}-{k},l_{q+1}-{k}\) (in the case \(\varrho >1\)) or \(m_{q}-{k},m_{q+1}-{k}\) (in the case \(\varrho =1\), \(\sigma >|p_{W}|^{-1}\)) is greater than one for q taken large enough. Hence, one can find a subsequence of terms of (IV2) which grows in modulus, and \(\tilde{c}_{{k}}\) diverges for all k. \(\square \)

Lemma E.1

(Coefficients \(\{\tilde{\alpha }_{k}\}\) for quotients of polynomials) Let \(g_{m}\) and \(h_{n}\) be polynomials of degrees \(m\) and \(n\), and consider

$$\begin{aligned} \textstyle f(u) = \frac{g_{m}(u)}{h_{n}(u)}, \quad \forall u\in \mathbb {R}_{{\ge 0}}. \end{aligned}$$

For \(\tau >0\), recall the polynomial \(p_{k}\) given in Table 2 and define \(f_{k}(s) = f(s) \, p_{k}({2s}/{\tau }-1)\) under the assumption \(\mathcal {P}(f_{k})\cap [0,\infty )=\emptyset \). The Fourier coefficients of \(f\) satisfy, for \(k\ge 0\),

$$\begin{aligned} \tilde{\alpha }_{k}= & {} \sqrt{\pi } \sum \nolimits _{q=0}^{l({k})} \frac{ \zeta _{-q}\, (-\tau )^{q} }{ {q}! \Upgamma \left( \frac{1}{2}-q\right) }\nonumber \\&-\, \sum \nolimits _{a\in \mathcal {P}(f_{k})} Res _{s=a}\left( { f_{k}(s) \, {s}^{-\frac{1}{2}}]_{-\pi } {(s-\tau )}^{-\frac{1}{2}}]_{-\pi } }\right) , \end{aligned}$$
(E9)

where \(l({k})=\max (0,m-n+k)\) is the largest nonnegative integer \(l\) such that \(\lim \nolimits _{s\rightarrow 0} s^{l} f_{k}\left( {1}/{s}\right) \) is finite, and \(\{\zeta _{q}\}\) are the coefficients of the Laurent series at \(+\infty \) of the analytic continuation of \(f_{k}\), i.e.,

$$\begin{aligned} \begin{array}{ll} \zeta _{q} = \frac{1}{{\left( l({k})+q\right) }!} \lim \nolimits _{s\rightarrow 0} \frac{d^{l({k})+q}}{ds^{l({k})+q}} \left[ s^{l({k})} f_{k}\left( \frac{1}{s}\right) \right] ,&\quad (q=-l({k}),\dots ,\infty ). \end{array} \end{aligned}$$
(E10)

See [9] for a derivation derivation of Lemma E.1 by the technique of contour integration.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bilenne, O. Dispatching to parallel servers. Queueing Syst 99, 199–230 (2021). https://doi.org/10.1007/s11134-021-09713-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11134-021-09713-y

Keywords

Mathematics Subject Classification

Navigation