Abstract

We consider two parallel queues, each with independent Poisson arrival rates, that are tended by a single server. The exponential server devotes all of its capacity to the longer of the queues. If both queues are of equal length, the server devotes of its capacity to the first queue and the remaining to the second. We obtain exact integral representations for the joint probability distribution of the number of customers in this two-node network. Then we evaluate this distribution in various asymptotic limits, such as large numbers of customers in either/both of the queues, light traffic where arrivals are infrequent, and heavy traffic where the system is nearly unstable.

1. Introduction

We consider a nonsymmetric version of the longer queue model. Here there are two parallel queues, each fed by a Poisson arrival stream. There is but a single server who tends to the longer of the two queues. If the number of customers in each queue is the same, then the server devotes of its capacity to the first queue and to the second queue, with . We let () denote the number of customers in the first (second) queue, the two arrival rates are and , and the server works at rate . Note that the total number of customers, , in the two-node network behaves as the standard model, so in the steady state we have the geometric distribution , where , assuming the stability condition . The “symmetric case” corresponds to (thus ) and , and this was analyzed in detail by Flatto [1].

Such models were proposed by Zheng and Zipkin [2] to study problems in inventory control. In [2] finite capacities were assumed in the two queues, and the authors studied numerically the steady state probabilities , in terms of the capacity size and also for different service disciplines, such as the longer queue (LQ) discipline here, and also the first-come-first-served discipline.

In [1] the author used two-dimensional generating functions and analyticity arguments and obtained explicit expressions for , in the symmetric case, as contour integrals. Then asymptotic results were derived for the joint distribution , as and/or becomes large, and also the marginal tails and various conditional limit laws were obtained. In this paper we generalize some of the results of [1] to the nonsymmetric model, and we will show that now many of the asymptotic results become quite different. As in [1] we assume that the model is preemptive, so if (and thus the server works on both queues) and a new arrival occurs to the first queue, then the server switches immediately all its capacity to the first queue. The more difficult nonpreemptive version of the LQ model was studied by Cohen [3], in the case of Poisson arrivals and general service times. The problem is reduced in [3], using generating functions, to a functional equation which is furthermore converted to a Riemann-Hilbert boundary value problem.

The present problem corresponds to a random walk in the quarter plane (as ), and other examples of such problems include shortest queue (SQ) problems [46], fork-join models [7, 8], two coupled processors with generalized processor sharing [911], and two coupled tandem queues [12]. General techniques for solving such problems are discussed in [1315] and they involve functional equations, analyticity arguments, and singular integral equation methods [16].

For the present model the analyticity arguments are fairly simple, and we focus mostly on the asymptotic properties of the solution. We will show that these asymptotics are quite different from those of the symmetric model in [1]. After obtaining exact integral representations for , and also the marginal probabilities and , we asymptotically evaluate these integrals for and/or large. We use standard techniques, such as the Laplace method, saddle point method, singular analysis, and the Euler-MacLaurin formula. Good general references on the asymptotic evaluation of integrals and sums are the books [1721].

In addition to this model being interesting on its own, many variants of shortest queue problems, such as ones with multiple servers and finite capacities, can be asymptotically reduced to LQ models of the type considered here (see [22, 23]). For example, in [23] we showed that the finite capacity version of the standard symmetric SQ model (analyzed in [4, 5]), where and is the capacity, asymptotically reduces to the symmetric LQ model in [1], if we consider the process , which measures the number of spots available in the two waiting rooms. Then having a thorough understanding of the nonsymmetric LQ model and its asymptotics will allow us to analyze, at least in some asymptotic limits, nonsymmetric variants of SQ models.

The remainder of this paper is organized as follows. In Section 2 we summarize all of the main results, both exact and asymptotic. They are listed in Theorems 15, and some discussion/interpretation appears following each theorem. In Section 3 we briefly derive the exact expressions for ; in Section 4 we derive asymptotic properties of for large. In Section 5 we derive light traffic (where ) and heavy traffic (where ) results.

2. Problem Statement and Summary of Results

We let be the numbers of customers in the two parallel queues, and let be the joint queue length distribution in the steady state. The two arrival rates are and , the exponential server works at rate , and . If the server works on the first (second) queue, but if the server works at rate on the first queue and rate on the second, with . The symmetric case corresponds to and . We henceforth assume the stability condition .

In Figure 1 we sketch the transition rates for the random walk , which illustrates the discontinuity along the “interface” . The main balance equations are We also have the following three interface equations: the two boundary conditions and the corner conditions The normalization is Note that the elementary difference equations (6) and (7) may be solved immediately to obtain and , up to multiplicative constants. In Section 3 we analyze (1)–(9) to obtain in the forms of contour integrals, which we summarize below.

Theorem 1. For , the steady state distribution is as follows: (i): where the integral is over a small loop about ,  (ii): (iii):

We next evaluate in various asymptotic limits, to gain more insight into the structure of the joint distribution. Writing to emphasize the dependence on the model parameters, we clearly have the symmetry relation Thus it is sufficient for the asymptotics to assume that and we also note that the expressions in Theorem 1 are consistent with (22).

We will show that the asymptotics are quite different whether , or . In Figure 2 we sketch the curve in the parameter plane, for . Note that the curve may also be written as and passes through the points and . We first give results for when , and note that this includes the symmetric case of Flatto [1].

Theorem 2. For and one lets be the queue length ratio. The following asymptotic expressions then hold:(i), (ii) with , where and can be computed from (12)–(14).(iii) with (thus ) (iv) with (v)(vi) with , (vii) with (thus ) (viii) with and can be computed from (14), (16), and (17).(ix),

Note that, in view of (12) and (16),

In Theorem 2 we listed the expansions of in order of decreasing , from in (25) to in (46). In the symmetric case and , and then and Theorem 2 is consistent with the results of Flatto in [1] (there the cases , and , were not considered, and the limits where were not treated in as much detail). Note that (34) and (38) correspond to “product form” approximations to . The expressions in (25) and (46) are actually exact when and , respectively, in view of our comments below (9). Our analysis in Section 4 will also indicate how to compute higher order correction terms in the various asymptotic series.

Next we take . Now some of the asymptotic results for will be very different, while in other ranges of they will be similar to those in Theorem 2. We emphasize below the formulas that are different.

Theorem 3. For and one has the following asymptotic expansions:(i) with and (ii)(iii) with (including the limit ), (iv) with (thus ) and is the parabolic cylinder function of order .(v) For and , (25) applies. For with , (27)–(30) hold. For with , (42)–(45) hold. For and , (46) applies.

Thus when the asymptotics of are different in eight ranges of , but the final result in four of these is the same as the case . Note however that if (27) applies for all , so the transition for is now absent. Also, if , (42) applies for rather than . In (57), is computed by setting , cf. (49), in (16), and we also note that .

The results in (49)–(60) are very different in form from the symmetric case asymptotics in [1]. For example, the diagonal probabilities in (52) contain the additional algebraic factor . In (57) the parabolic cylinder function can be computed, for example, from where is a vertical contour in the complex -plane with , and we have for the correction term in (57). Also, and for we can express in terms of a Hermite polynomial. In (57) we give a two-term asymptotic approximation in the transition range . Also, it may be shown that the leading term in (57) is just the limiting form of (42), as . The expression in (55) applies both to with , and to with . In the latter case we can approximate the factor by . The parabolic cylinder function has the asymptotic behaviors so as in (57) (corresponding to ) the correction term begins to dominate the leading term (which occurs for or ). Then we can show that the expression in (57) for agrees with (55) for , and hence the two expansions asymptotically match. Note also that the scaling in (57) has .

Next we study the transition range in parameter space. This will lead to a new set of asymptotic results which will show, for example, how the formula in (52) for changes to the purely geometric approximation in (35). To quantify the closeness to the curve in Figure 2 we write where and can have either sign. If we are exactly on the transition curve. For small , in certain ranges of , the results in Theorems 2 and 3 still apply. For example, if and or and Theorems 2 and 3 agree, and then no transition range result is needed. As , in (26), and then (27) will hold for all in the transition case. Thus for the transition case will require a new asymptotic result only if . For the asymptotic result in (42) will apply for , where now, since , (cf. (37) and (54)). Then if , we will need different asymptotic results only when , including and , where Note that is the limit of both and , if we replace by , as (37) leads to (setting , ) since corresponds to . Similarly, (54) leads to in this limit.

Since we will now have the state variables and large, and also small, it is necessary to relate these. In Theorem 4 we summarize the transition case results, scaling and in terms of .

Theorem 4. For and with , one has the following, where and .(i), (ii)(iii)(iv), (thus ) and is the parabolic cylinder function of order   (see (61)).(v), , and is defined by the contour integral where is to the right of all singularities of the integrand, including the pole at if .

If then and the term with the integral in (66) is absent. In Theorem 4 we wrote most results in terms of and . As it can be shown that (66) asymptotically matches to (35), and as , (66) matches to (52). Similarly as (68) matches to (27) while as (68) matches to (48). Some of these asymptotic matching relations are further discussed in Section 4. In (70), depends on and also on the ratio , in view of (71). The results in Theorem 4 show that for it is natural to scale , and then either independently, or . The expression in (74) is similar to, but somewhat more complicated than, that in (72). The function in (76) reduces to in the limit , and thus (72) may be viewed as a special case of (74)–(76). Note also that the results in Theorem 4 can be rewritten without introducing and in (63), as for example and is proportional to , and may be written in terms of the original parameters .

Next we consider some different asymptotic limits, those of “light” and “heavy” traffic. Light traffic corresponds to infrequent arrivals, where . Heavy traffic corresponds to nearly unstable systems, where . It turns out that the present model has two possible heavy traffic limits. In the first, which we call HTL1, we have with a fixed . Then most of the probability mass will occur in the range where and are large, but with . More precisely, if then and must be scaled to be , but with the difference fixed. In the second heavy traffic limit (HTL2) we again set but now let , with . Now the probability mass will become more spread out, with appreciable mass anywhere in the range , where and . The light and heavy traffic results are summarized below as Theorem 5.

Theorem 5. Light traffic: For one has
Heavy traffic limit with ,
Heavy traffic limit with , ,  Here is a vertical contour in the -plane, which lies to the right of all singularities.

The expression in (78) applies for , , and , and in the light traffic limit the discontinuity of along the diagonal will appear only in the higher order terms. In HTL1, (79)–(81) show a piecewise geometric distribution in the variable, and an exponential density in . For HTL2, writing (82) as we can easily show that so that to leading order the probability mass concentrates where with . From (84) we have but the total mass along the main diagonal is , which is smaller than the mass in (82). Then also with total mass , which is comparable to that along the main diagonal. The diagonals with have mass , which is smaller still. The integrands in (82), (84), and (86) have branch points at , are analytic at , and may have poles at , if .

This completes our summarization of the exact and asymptotic results. Despite the seeming complexity and the many separate cases, all the results follow from fairly standard asymptotic evaluations of the integrals in Theorems 1, as we will show in Sections 4 and 5.

3. The Exact Solution

We solve the difference equation(s) in (1)–(9) and thus obtain Theorem 1. We begin by introducing the three generating functions: and we note that , and Then (9) shows that

From (1) and (3) we obtain, after some calculation, and (2) and (4) lead to Equation (5) along then leads to the following relation between , , and :

If we define then from (96) and (88)–(90) we find that Using (93) with , (94) with and (95) with , we obtain for the simpler equation so that and then by (92), and thus . Then from (96) we have so that follows a geometric distribution with parameter , and hence behaves as the standard model.

The coefficient of in (93) has roots at while that of in (94) has roots at By using (95) to express in terms of and , eliminating in the right sides of (93) and (94), and requiring that be analytic at and that be analytic at , we obtain two equations for and : Solving the algebraic system in (98) and (99), and then using the result in (95) to compute , we hence obtain explicitly the right sides of (93) and (94). In particular, where is as in (12)–(14). Comparing (100) to (88) we conclude that for . Setting and inverting the generating function in (101) leads to (10) for . Similarly, inverting the double generating function for in (89) leads to (15), and we note that (10) and (15) are consistent with the symmetry . Finally, in (19) is obtained by inverting the generating function . We have thus established Theorem 1.

4. Asymptotics of the Joint Distribution

We derive Theorems 24 by expanding asymptotically the integrals in Theorem 1. We will use a combination of the saddle point method and singularity analysis. Good general references on techniques for asymptotically evaluating integrals can be found in [1721].

We need to understand the singularities of the integrands in (10), (15), and (19). There are clearly branch points where and , with Since we have and for the branch point at is farther from the origin than the one at . In fact, will never play a role in the asymptotics. The integrands are also singular at , where (10) has a pole of order , and (15) has a pole of order . The only possible other singular points are at the zeros of . We can easily verify that is a simple zero of all four functions , , , and , so all the integrands are analytic at . In the appendix we study in detail the algebraic equation , and show that the only possible zero is at but only if (if then has a branch point at ). If then the two branch points are the only singularities of the functions , , and in (12), (16), and (19). If , which is clearly true in the symmetric case, then is a simple pole of these functions (since ). In view of (105) and (106), we have , if the pole is present and the stability condition holds.

Consider first (19). If the pole at determines the asymptotic behavior, as it is closer to than the branch point at . Hence as Straightforward computations, using (14) and (21), show that Using (107)–(109) leads to (35), with (36).

Next we let simultaneously, setting , still taking , and first consider , where (10) applies. The integrand in (10) again has singularities at , , and . But now we write and let . Then with (110), (10) has saddle points where , or Solving (111) using (11), after some algebra we are led to the saddle in (28). There is also a second saddle at , where corresponds to replacing in (28) by . By evaluating and we can easily show that the directions of steepest descent at the saddle(s) are and . From (28) we have , so as the saddle approaches a branch point. Also, as . It is possible for the pole and saddle to coalesce. Setting and solving this equation for we find, again after some algebra, that where is given by (26). If , we have and if , we have . From (28) we can show that so that the saddle location decreases with the queue length ratio . If we deform the contour in (10), which is a small counterclockwise loop about , into a saddle point contour, on which . Then the new contour traverses the saddle in the direction(s) of steepest descent and the standard Laplace estimate of (10) is But and, in view of (111), Then using (11) we commute to find that is the same as in (30). Thus (113) is the same as the right side of (27), and we have established the asymptotics of for .

If and in deforming the loop in (10) to the saddle point contour we must take into account the contribution from the residue at the pole . But we have with equality only if , when . Thus the pole contribution dominates the saddle contribution and we have From (13) we find that and from (11) when . With (115)–(117) and (108) we have established (34) for , where gives the product form expression and corresponds to the multiplicative constant .

As we have and we must then reconsider the asymptotics of (10). From (28) as so we scale in the integral in (10) and consider the limit with . Then and (10) becomes asymptotically Since we see that (119) is the same as (25).

Now consider , where the saddle and pole are close to each other. This is a standard problem that is discussed, for example, in [17, 18]. We now expand the integrand in (10) about , ultimately scaling , and then the integrand will approach a limiting value as . We have, by Taylor/Laurent series, From (11) and (122) we find that since, at , . It follows that where was defined above (31). Also, from (11) we find that Then setting and scaling as in (32), (121) becomes Then we use (120) and (126) to get where is a vertical contour with . To obtain (127) we shifted the original contour in (10) into the circle , , and note that implies to leading order that . The integral in (127) can be expressed in terms of a parabolic cylinder function of order (see (61)), which can be expressed in terms of the standard error function, using the identity With (128) and (117), (127) becomes the same as (31), so we have derived the leading term for the range . We have thus covered all of the necessary ranges that have , and established items (i)–(v) in Theorem 2.

Now we consider ranges with where is given by the integral in (15). The analysis is completely analogous to the expansion of (10), so we merely sketch the details. The function in (16) has a simple pole at if , with residue where We write where Thus for simultaneously large will have a saddle where , and this leads to , as in (43). The saddle and pole coalesce where in (37). For we have and the pole determines the asymptotic behavior of . Then and this leads to (38) with (39), as now . For the saddle determines the asymptotics, and the estimate leads to (42), with (43)–(45). As the saddle and the result in (46) can be obtained by approximating the integrand in (15) for , similarly as in (118) and (119) (with replaced by , by , and by ). Thus we obtain (46). Finally, when the pole and saddle are close. Then we scale and expand the integrand in (15) near , with . This leads ultimately to (40) with (41). We have thus sketched the derivation of items (vi)–(ix) in Theorem 2.

Next we take and establish Theorem 3. Now the pole at is absent. We first consider the diagonal probabilities in (19). The only singularities of as the branch points at and , and the former determines the large asymptotics, as in view of (105). Expanding about will transfer immediately to an asymptotic series for , but here we focus on the leading term. From (14) and (21) we have where Note that . Using (133)–(137) we have

The expansion of will be in powers of , but only the odd powers will contribute to the asymptotics. Now, where we used the binomial expansion of and Stirling's formula. It follows that the leading term for is, in view of (138) and (139), with a correction that is relative to the leading term, which may be computed from the term in (138), and a refined Stirling approximation of the factorials in (139). Some of the algebra in our calculations is simplified by introducing and . Then in (136) factors as and if . After some calculation we find that so that and then is the same as the constant in (53). We have thus established (52).

Next we consider . The saddle point calculation that led to (27) does not depend on whether . If the pole at is absent, then (27) will apply for all . For and (25) again holds, since this calculation is independent of whether exists or not. However, now a different analysis is needed for , and the appropriate scale is to fix and let . As we have and then The expansion in (27) breaks down as (if ) as (30) shows that and thus the factor in (27) vanishes linearly as , which indicates a problem in the asymptotics.

We thus reexamine (10) for and large. We again employ singularity analysis and expand the integrand about the branch point . We let since in view of (13). Then where and we again used some algebraic factorization in the variables. From (11), and hence By multiplying (146) by (148) and using (139), the leading term for becomes But (149) is the same as (48), with and , since in (137) may also be written as We have thus established (48).

Now consider with . The integrand in (15) again has a saddle at , and as . For and we again obtain (46). As the saddle . Thus there is a critical value of , which we call , such that Using (130), and the facts that and , we obtain which is a linear equation for whose solution is given by (54). If then and the saddle point approximation in (42) holds. If the branch point at determines the asymptotics of , and we again use singularity analysis. Now we expand about , writing (17) as where If we furthermore scale , then It follows, by using (155)–(157) in (15), that where on . After some calculation we find that, again using and , The integral in (158) can then be evaluated as Combining (141) with (158)–(162) we obtain precisely the result in (54)–(56).

Now consider the case , where the branch point at is close to the saddle at . The standard scaling for such situations with coalescing singular points (see [18]) is and . By expanding in Taylor series about , setting and scaling as in (59), and letting we find that Using (163) and (164) in (15) and again expanding as in (157) we obtain where is to the left of all singularities. Here we also used the definition of in (58). The integral(s) in (165) may be easily evaluated using where the latter follows from (61), after replacing by . But by (157), and then (165) with (166) give precisely the approximation in (57)–(60). Note that is the same as .

We can easily compute higher order terms in the expansion, and our analysis shows that the asymptotic series will now involve powers of . Actually, the leading term in (57), which has a Gaussian dependence on (hence on ), can be obtained by simply expanding the saddle point approximation in (42), for . However, the correction term is necessary to see the transition to the range , where (55) applies. In view of (62) the correction term becomes comparable to the leading term when or, roughly, when . Thus for or with the Gaussian dominates while for the term dominates. Note that and for sufficiently negative, by (62). Thus for both the and terms in (57) are positive. The asymptotic matching of (57) for (with ) with (55) as is then easily verified, as the approximation in (55) has an algebraic singularity as , which agrees with the second formula in (62). Also, up to a multiplicative constant agrees with . This completes the derivation of Theorem 3, where .

Next we analyze how the results in Theorem 2 transition to those in Theorem 3, as decreases through . We could simply assume that and then obtain the necessary asymptotic results. But to see the transition it is necessary to also analyze cases where . To make this more precise we write as in (63), and assume that . Then according as . Since only the product is important, we can set , according to the cases . Then we must relate the small parameter to the large parameters , and we show below that a natural scaling is to take as .

The asymptotic results for and are the same in Theorems 2 and 3, and thus no transition is needed here. We can write these results in terms of, say, rather than and expand for small to somewhat simplify the expression in (25), but we will not do so here. If the saddle point approximation in (27) applies for all , while if it applies only for . But (cf. (26)) as so it will apply for any fixed . But will require a separate analysis. We also note that the sector , where the product form solution in (34) applies, shrinks to zero. Thus if such an approximation will play a role here, it must be contained near .

We begin by considering the diagonal probabilities , using the scaling in (168). We will approximate in (19) in such a way that the integrand approaches a nondegenerate limit. The pole at , if it exists, is now close to the branch point , since . Note that if . Introducing the scaling , so that , we have . Now we write near as From (134) we find that while (136) or (141) shows that since and thus . From (137) or (151) we also find that as Then from (169)–(172) we see that becomes comparable to if , or . Then setting and using (169)–(172), we can approximate (19) to obtain Thus it is useful to introduce and then we evaluate the contour integral in (173) as, changing , Here we used a conformal map , some contour deformations, (61) with , and an identity that relates to the standard error function, which yields the last equality in (174). Using (174) in (173) and noting that , we obtain (66).

Now consider and we already discussed the case with the transition range scaling in (168). We expand now (10) similarly as we expanded (15). For , the factor can be approximated by for . We thus let so that . Also, as , and then Hence (10) becomes Now and can be evaluated as in (171) and (172), and, as , in (147) becomes Then (177) can be evaluated similarly to (173), and we ultimately obtain (68), with as in (69).

For we need only consider the ranges and , where is in (64), with being the limit of both and , as or . For the saddle at exceeds the branch point and hence the latter determines the asymptotic behavior of . For a fixed , we scale and use We again expand in the form in (169), with now replaced by , where Using (161) with replaced by , along with (180) and (181), the integral in (7) becomes Here we let and used (171) and (172) to approximate and . Scaling as in (178) and evaluating the integral in (182) similarly to (173) leads to (70). If , the same analysis applies, as then we can simply replace by 1 and by in (70) and (71), but must maintain the factor .

When we let . Now the saddle will be close to the branch point (with also by (168)). We scale with and introduce as in (73). For we also have and then Since and we have and then Expressing the integral as a parabolic cylinder function of order (see (61)) we obtain the expression in (72), with (73) or (185).

If we let but consider even larger values of , with , a slightly different expansion applies. Now both and in (186) become . Setting , we use in (15) to ultimately obtain the expression in (74), which involves the contour integral in (76). The function can be expressed as an infinite sum of parabolic cylinder functions, as

This completes the analysis of the transition range where , and we have thus established Theorem 4.

5. Light and Heavy Traffic

We establish Theorem 5. First consider the light traffic limit, where . We now scale and then, from (11), (12), (16), and (20), obtain Hence the diagonal probabilities have the limiting form where we used the binomial theorem, and denotes the coefficient of in the Taylor expansion of what follows.

For we similarly obtain and this holds also for . By refining (191) with (190) we can obtain higher order terms, as an expansion involving powers of and . We have thus established (78).

For the first heavy traffic limit (HTL1) we let , replace by and also set Then after some calculation we find from (14) that and from (21), (13), and (17) that

Hence the limiting form of (19) is which is the same as (80). With the scaling the loop integral in (19) can be approximated by the vertical contour , on which in this case. Since and we use (200), (195), (197), and (198) to approximate the integrands in (10) and (15), and this leads to (79) and (81).

We can also derive the HTL1 limits directly from the asymptotic formulas in (34), (35), and (38), as for and . This shows that it is permissible in this case to first let with a fixed and then let . Note also that in HTL1 the condition is certainly satisfied, so Theorem 2 applies.

In the second heavy traffic limit, HTL2, we again let and , but now is small. We now set and , and we have Then after some calculation we find that as and thus is and given by the expression in (84) (again we have ). Since as , the right side of (84) has probability mass along , and we can write The presence of the term in (205) indicates that different approximations are needed for for and for small , such as . But in HTL2 there is little mass along the diagonal as a whole.

In HTL2 we also obtain so that , as in (83), and then is given by (82). We also note that, in view of (203) and the scaling ,

If is the limit of , for , in HTL2, then by contour integration so that is a proper two-dimensional density function, with support in the range . In (208), on .

For in HTL2 we have so that . For we also have and Then as below (86). Thus for we have derived (86), and this completes the analysis of HTL2. Note that in HTL2 both and are .

Appendix

Here we study the roots of in (14). We have and so that under the stability condition . Setting we have Now when and when (since ). We thus have real for real and . From (A.4) we conclude that is a convex function of for . This is obvious for and for we note that and , in view of (A.2). Also, (A.3) shows that and . In view of the convexity and the fact that and we conclude that has at most two real roots. But as we discussed in Section 4, is a root of if . Setting yields and setting leads to To determine the sign of we use (49) and the variables and . Then is equivalent to which simplifies to so that according as or , that is, when is a root. It follows that if , and are two roots of , and if , is the only root. Since and there can be no roots for .

Now we consider the possibility of having complex roots of . Writing we eliminate the radicals to obtain Thus if is a root of it must be a root of (A.10). After some simplification and factoring using MAPLE, we find that solving (A.10) is equivalent to solving the quartic equation where Here , , have different meanings from those in Section 4. Clearly , , are all positive. If , and then is a root (A.11), but it cannot be a root of .

If , has no roots where , but we can show that it has two real roots in the range . Letting and we must show that the discriminant . After some calculations we find that where But , since . Thus for and the last factor in (A.11) has two distinct negative roots. If or , there is a double root in the range , as then . In either case these negative roots cannot be roots of .

We have shown that any root of must be a root of (A.11), and this quartic has only real roots. Then only one and can be roots of , and we already showed that the former is always a root while the latter is a root if and only if . Note that so that for , is always positive, while is proportional to , which vanishes along the transition curve.

Acknowledgments

Chalres Knessl work was partly supported by NSA Grant H 98230-11-1-0184. Haishen Yao was supported by PSC-CUNY research award no. 64349-0042.