Most hummingbirds and some species of nectar bats hover while feeding on floral nectar. While doing so, they achieve some of the highest mass-specific values among vertebrates. This is made possible by enhanced functional capacities of various elements of the ‘O2 transport cascade’, the pathway of O2 from the external environment to muscle mitochondria. Fasted hummingbirds and nectar bats fly with respiratory quotients (RQs; ) of ∼0.7, indicating that fat fuels flight in the fasted state. During repeated hover-feeding on dietary sugar, RQ values progressively climb to ∼1.0, indicating a shift from fat to carbohydrate oxidation. Stable carbon isotope experiments reveal that recently ingested sugar directly fuels ∼80 and 95% of energy metabolism in hover-feeding nectar bats and hummingbirds, respectively. We name the pathway of carbon flux from flowers, through digestive and cardiovascular systems, muscle membranes and into mitochondria the ‘sugar oxidation cascade’. O2 and sugar oxidation cascades operate in parallel and converge in muscle mitochondria. Foraging behavior that favours the oxidation of dietary sugar avoids the inefficiency of synthesizing fat from sugar and breaking down fat to fuel foraging. Sugar oxidation yields a higher P/O ratio (ATP made per O atom consumed) than fat oxidation, thus requiring lower hovering per unit mass. We propose that dietary sugar is a premium fuel for flight in nectarivorous, flying animals.

Metabolic fuel use during exercise is a subject of great interest to human exercise physiologists, comparative biochemists and ecological physiologists. With few known exceptions, e.g. tsetse flies (Bursell and Slack, 1976), a pattern that has emerged is that energy metabolism during exercise relies primarily on carbohydrate or fat as substrates. Another pattern, based on studies of humans and several other vertebrate species, is that as exercise intensity increases and the maximum rate of O2 consumption () is approached, the fractional contribution of fatty acid oxidation to ATP production declines whereas that of carbohydrate oxidation increases (Brooks, 1998; Weber and Haman, 2004). As animals approach during exercise, muscle glycogenolysis accounts for an increasing fraction of the carbohydrate used (Weber et al., 1996). The greater reliance on glycogen than on blood glucose in animals exercising close to VO2max appears to be due to limitations to glucose phosphorylation in muscle fibers (Fueger et al., 2004). However, the capacity to fuel exercise metabolism directly from dietary sources is also limited. For example, in humans, only ∼30%, at most, of exercise metabolism can be directly fueled by ingested sugar (Jentjens et al., 2004). In this paper, we describe remarkable cases of deviation from the above paradigm, made possible through the operation of a pathway we name the ‘sugar oxidation cascade’.

The sugar oxidation cascade is the path of carbon from flowers, through digestive and cardiovascular systems, across capillary walls, muscle cell membranes and into the mitochondria in hovering, nectarivorous vertebrates (Fig. 1). The operation of this pathway enables hummingbirds and nectar bats to engage in aerial refueling, i.e. to use recently ingested sugar to directly fuel their exercising muscles during hovering flight. In many respects, the sugar oxidation cascade is analogous to the ‘oxygen transport cascade’ (Weibel, 1984; Weibel et al., 1981), which is the path of O2 from the external environment, through the respiratory and cardiovascular systems and into the mitochondria of exercising muscles. The two cascades share some common elements, operate in parallel and ultimately converge in the mitochondria of the flight muscles (Fig. 1). Although we shall focus on our own work using hummingbirds (primarily rufous hummingbirds, Selasphorus rufus Gmelin 1788) and, more recently, Pallas' long-tongued bats (Glossophaga soricina Pallas 1766; hereafter nectar bats), we shall also draw on the findings of others to describe the elements of the cascade and the biological context in which it operates.

While foraging for free coffee in the 1970s, Peter Hochachka and his graduate students pondered a metabolic mystery: when hummingbirds hover to forage on floral nectar, they ingest mainly sugar. Do they immediately synthesize fat from the ingested sugar, only to break it down shortly after to fuel foraging flight? At this time, the remarkable ability of birds to build up fat stores and to use it to fuel energy metabolism was already well appreciated (Blem, 1976). Although hummingbirds, like many other species of birds, were known to undergo premigratory fattening and to use fat to fuel migratory flight (Odum et al., 1961), it seemed implausible, even to naïve biochemists, that they would use fat to fuel foraging activity.

Much has been learned concerning hummingbird biology in the decades that followed. Rufous hummingbirds spend about half of their lives migrating between breeding grounds in the Pacific Northwest and overwintering habitats in Mexico (Calder, 1987). As fat stores are depleted, they stop along their migration route to refuel. Hummingbirds obtain most of their dietary calories from floral nectar (Powers and Nagy, 1988). They visit flowers that produce nectars rich in sucrose (Baker et al., 1998) and prefer sucrose solutions to solutions of glucose + fructose (Martinez del Rio, 1990). Fat synthesis from ingested sugar occurs at rates that allow the birds to gain up to 10% of body mass per day (Carpenter et al., 1983). Rufous hummingbirds tend to fly for short durations while foraging on floral nectar (Diamond et al., 1986). Foraging often occurs in subalpine meadows where, in the early morning hours, ambient temperature (Ta) can be near freezing (C. L. Gass, personal communication) (Gass et al., 1999). At low Ta, thermogenic mechanisms are activated (Bicudo et al., 2002), resulting in elevation of rates of energy expenditure (Lasiewski, 1963; Lopez-Calleja and Bozinovic, 1995; Lotz et al., 2003). To maintain energy balance or to achieve net energy gain and synthesize fat under these conditions, the birds increase foraging activity (Gass et al., 1999; Suarez and Gass, 2002). Low air densities encountered at high altitude further raise the energetic costs incurred during foraging (Welch and Suarez, 2008). Thus, hummingbird foraging often occurs under environmental conditions easily characterized as ‘extreme’.

In contrast with hummingbird-visited flowers, those visited by nectarivorous bats produce nectars with low sucrose and high glucose + fructose concentrations (Baker et al., 1998). However, the nectar bats that we recently studied do not discriminate between solutions of sucrose and glucose + fructose of equal energy content (Rodriguez-Pena et al., 2007). Glossophaga soricina does not migrate, unlike some other nectarivorous bat species (Fleming et al., 1993; Morales-Garza et al., 2007), and does not accumulate much body fat (McNab, 1976). However, they are capable of energetically expensive hovering (Voigt and Winter, 1999; Winter et al., 1998), allowing mechanistic comparison of a nectarivorous hovering mammal with hummingbirds.

Hummingbirds in flight sustain some of the highest mass-specific rates of aerobic metabolism (, where Mb is body mass) known amongst vertebrates (Suarez, 1992). Nectar bats display lower but, nevertheless, impressive hovering values (Voigt and Winter, 1999; Winter et al., 1998), similar to the maximal rates observed in shrews exposed to low Ta (Fons and Sicart, 1976). Respiratory physiologists have found that the flux of O2 is regulated at multiple steps in the oxygen transport cascade (di Prampero, 1985; Jones, 1998; Wagner, 1996). It is probably because of such distributed regulation (i.e. the absence of a single, rate-limiting step) that enhanced functional capacities at multiple steps in the pathway of O2 have evolved in hummingbirds and bats. These have been the subject of previous reviews (Maina, 2000; Suarez, 1992; Suarez, 1998) and include high lung O2 diffusing capacities, large hearts and high heart rates, high cardiac outputs, high haematocrits, high muscle capillary densities, high mitochondrial volume and cristae surface densities.

Both hummingbirds and nectar bats possess flight muscles consisting exclusively of fast-twitch, oxidative fibers (Grinyer and George, 1969; Hermanson et al., 1998; Suarez et al., 1991). It is likely that, when in flight, 90% or more of their whole-body and (rate of CO2 production) values are accounted for by flight muscles (Suarez, 1992; Taylor, 1987). Given their small body masses and high values during flight, whole-body O2 and CO2 fluxes measured by respirometry yield respiratory exchange ratios () that are likely to approximate cellular respiratory quotients (RQs). These data can be used to determine the nature of metabolic fuel(s) oxidized and to estimate ATP turnover rates, given what is known concerning the stoichiometries of carbohydrate and fatty acid oxidation, as well as mitochondrial oxidative phosphorylation (Brand, 2005).

We have used mask respirometry, essentially as previously described (Bartholomew and Lighton, 1986), to measure and of hummingbirds and nectar bats as they hover to feed on sugar solutions dispensed within the mask. When these animals begin to forage in the fasted state, RQ values are close to 0.7, indicating that fatty acids are oxidized to provide most of the energy for flight (Suarez et al., 1990; Welch et al., 2008) (Fig. 2). As further feeding occurs, RQ values progressively increase to ∼1.0. This indicates that, in both hummingbirds and nectar bats, carbohydrate progressively takes over as the main fuel for oxidative metabolism in the flight muscles as they hover to feed on sucrose. The data shown in Fig. 2 were obtained from rufous hummingbirds and nectar bats; similar results have been obtained from Anna's (Calypte anna) (Welch et al., 2008) and broadtailed (Selasphorus platycercus) (Welch et al., 2006) hummingbirds.

Further insights into fuel use during foraging come from the use of carbon stable isotopes (Welch et al., 2006; Welch et al., 2008; Welch and Suarez, 2007). These experiments take advantage of the lower 13C/12C ratio of sucrose synthesized by C3 photosynthesis in beets compared with that of sucrose from sugar cane, which is made by C4 photosynthesis. This results in a more negative (i.e. more 13C-depleted) δ13C of beet sugar than cane sugar, where:
formula
(1)
and Rstd=[13C]/[12C] of marine limestone from the Pee Dee Cretaceous belemnite in South Carolina or an artificial version from Vienna (see McNevin et al., 2007). Hummingbirds and nectar bats maintained on diets enriched in beet sugar expire CO2 with δ13C values similar to that of beet sugar. In the fasted state, initial foraging flights yield RQ values close to 0.7, indicating fatty acid oxidation. However, the fatty acids oxidized in these animals were previously synthesized from beet sugar, so δ13C values of CO2 expired during hovering are low. The animals were then provided cane sugar via the feeder in the mask. Repeated feeding bouts were accompanied by progressively increasing δ13C values of expired CO2. Fig. 3 shows that the δ13C values increase with RQ, indicating that the direct oxidation of recently ingested cane sugar accounts for an increasing fraction of the energy required for flight as the animals transition from oxidizing fat to oxidizing carbohydrate. When the animals are repeatedly foraging and consistently display RQ values close to 1.0, dietary sugar oxidation provides ∼95 and 80% of the energy required for flight in hummingbirds and nectar bats, respectively (Fig. 4). Reliance of oxidative metabolism primarily on recently ingested sugar was also previously shown in nectar bats (Voigt and Speakman, 2007) using breath samples from animals restrained shortly after exercise. By contrast, ingested sugar can fuel only ∼30%, at most, of human energy expenditure during exercise (Jentjens et al., 2004).

The operation of the sugar oxidation cascade is analogous to the aerial refueling performed by certain types of high-performance aircraft in that the ingested fuel is quickly used by oxidative reactions that convert the energy stored in organic compounds into mechanical work and heat. In common with such aircraft, nectarivorous, flying vertebrates require high capacities for the delivery of O2 and fuel to their flight motors. Evidence for this is seen at the level of the digestive system: enhanced digestive capacities are made possible by high intestinal sucrase activities (Hernandez and Martinez del Rio, 1992; Schondube et al., 2001) as well as the combined use of active and passive mechanisms (involving paracellular movement) for sugar movement across the intestinal epithelium (McWhorter et al., 2006). The use of both active and passive mechanisms by fruit bats (Caviedes-Vidal et al., 2008) suggests that both mechanisms occur in nectar bats as well. Hummingbird cardiac outputs are estimated at approximately 5 times body mass per minute (Johansen, 1987), ensuring high capacities for convective transport of blood-borne fuels, whereas high capillary to muscle fiber surface area ratios (Mathieu-Costello et al., 1992; Suarez et al., 1991) would enhance diffusive capacities, not just for O2 and CO2, but for metabolic fuels as well.

Upon entry into exercising muscle fibers, glucose is phosphorylated to glucose 6-phosphate (G6P). Nectar bat and hummingbird flight muscles possess extraordinarily high capacities for glucose phosphorylation to G6P, catalyzed by the enzyme hexokinase (Suarez et al., 1990; Suarez et al., 2009). These biochemical capacities are estimated by measurement of enzyme maximum velocity (Vmax) values (where Vmax=kcat×[E], kcat is catalytic efficiency and [E] is enzyme concentration); these establish the upper limits to physiological flux rates (Newsholme and Crabtree, 1986; Suarez, 1996). Given high hexokinase Vmax values, it seems likely that flight muscle membranes in hummingbirds and nectar bats would also possess high capacities for glucose transport, made possible by high levels of glucose transporter (GLUT) expression. No information is yet available concerning hummingbird GLUTs. However, in sparrow (Passer domesticus) skeletal muscles, Sweazea and Braun report the presence of mRNA coding for GLUT1 and GLUT3 and the absence of GLUT4 (Sweazea and Braun, 2006). In addition, they show immunohistochemical evidence of protein expression of GLUT1 and GLUT3, as well as western blots showing the absence of GLUT4. Similarly, Seki et al. report the absence of GLUT4 in broiler chickens (Seki et al., 2003). Sweazea and Braun argue that previous reports of GLUT4 in avian skeletal muscles are probably erroneous (Sweazea and Braun, 2006). On the basis of the established role of GLUT4 in exercise and insulin-stimulated glucose transport in mammalian skeletal muscles (Huang and Czech, 2007), nectar bats would be expected to possess high levels of GLUT4 protein expression in their flight muscles. This appears to be the case, based on preliminary results (R. Lee-Young, D. Wasserman and R.S., unpublished).

In exercising hummingbird and nectar bat muscle fibers, abundant mitochondria operate as O2 sinks and intracellular gradients drive diffusive O2 fluxes from capillary red blood cells to mitochondrial cytochrome c oxidase. The O2 and sugar oxidation cascades converge at the level of flight muscle mitochondria, where the oxidation of each mole of C6H12O6, requiring 6 mol O2, leads to the production of 6 mol CO2 by decarboxylation reactions in the Krebs cycle and 6 mol H2O by the cytochrome c oxidase reaction, as well as 2.41 mol ATP per O atom consumed (Brand, 2005), assuming operation of the malate–aspartate shuttle for cytoplasmic redox balance (Suarez et al., 1986; Suarez et al., 1990). It appears that the evolution of high capacities for flux through the O2 transport cascade, required because of small body size and high mass-specific power output during flight, has partly set the stage for high capacities for carbon flux through the sugar oxidation cascade as a consequence of multiple shared elements.

According to Chantler: “The most noble aim of the biochemist, often discussed when inebriate, seldom when sober, is to relate the in vitro to the in vivo” (Chantler, 1982). It is therefore of interest to consider the rate at which the sugar oxidation cascade might operate during hovering flight. To address this, it is useful to first consider the more general question of how energy metabolism is regulated in muscles performing steady-state work. Muscles are biological machines and their mechanical power output largely determines their rates of ATP hydrolysis. In locomotory muscles performing repeated cycles of contraction and relaxation during steady-state exercise, the power output of a given volume of muscle is a function of its operating frequency (contraction cycles/time), stress (force/cross sectional area) and strain (fractional change in length per contraction) (Pennycuick and Rezende, 1984). In synchronous muscles, most of the ATP used during exercise is hydrolyzed by actomyosin-ATPase and Ca2+-ATPase (Homsher, 1987; Homsher and Kean, 1978; Szentesi et al., 2001). Metabolic control analysis performed using skinned rat soleus fibers reveals that, among various processes, the greatest degree of control (i.e. the highest control coefficient) over mitochondrial respiration is exerted by ATP hydrolysis (Wisniewski et al., 1995). Given current knowledge concerning the stoichiometries of fuel oxidation and oxidative phosphorylation (Brand, 2005), values during hover-feeding (when RQ=1) can be used to estimate rates of muscle ATP turnover and flux rates through the sugar oxidation cascade (Table 1). Interspecific comparisons reveal that hexokinase operates at much higher fractional velocities in the flight muscles of nectarivorous animals (Suarez et al., 1990; Suarez et al., 2009) than in the locomotory muscles of other species (Suarez et al., 1997). Thus, in rufous hummingbirds and nectar bats, high rates of glucose phosphorylation in vivo result from high levels of hexokinase expression as well as the operation of this enzyme at high fractional velocities during foraging flight (Suarez et al., 1990; Suarez et al., 1997; Suarez et al., 2009).

The impressive rate at which the sugar oxidation cascade can operate is dramatically illustrated by the results of laboratory experiments simulating the cold mornings encountered by migratory rufous hummingbirds at foraging sites in the late summer (Gass et al., 1999). These experiments were performed in an environment chamber wherein Ta was held at 5°C. Because digestive efficiency in hummingbirds is close to 100% (Diamond et al., 1986; McWhorter and Martinez del Rio, 2000), the energy intake rate is calculated from the volume and concentration of sucrose solution ingested over the 4 h duration of the experiments. Mass gain, mainly in the form of fat, results when energy intake rates exceed rates of energy expenditure. Mass loss indicates that intake rates are insufficient to meet energy requirements and depletion of fat stores results. Maintenance of mass indicates that the time-averaged dietary energy intake rate equals energy expenditure. In these experiments, low Ta combined with low energy content of sucrose solutions dispensed at the feeder elevated energetic costs and drove up foraging activity. Fig. 5 shows that hummingbirds fed various volumes of 15 and 20% sucrose at 5°C tended to lose mass, whereas maintenance of mass or mass gain were possible when birds were provided 30% sucrose. The regression intersects the line, indicating zero mass change at a maximum sustained metabolic rate (Hammond and Diamond, 1997; Peterson et al., 1990) of 1 W. When energy metabolism is completely fueled by dietary sucrose, the time-averaged rate of glucose ingestion and oxidation by a 4 g bird, given an energy content of 16.5 kJ g–1 sucrose, is 218 mg h–1 (Gass et al., 1999). Time-averaged metabolic rates of ∼0.81 W are sustainable for 12 h periods (Beuchat et al., 1979). Thus, at low Ta, rufous hummingbirds achieve the highest known maximum sustained metabolic rates among vertebrates, fueled by the sugar oxidation cascade when dietary energy intake rates equal rates of energy expenditure (Hammond and Diamond, 1997; Peterson et al., 1990).

The scientific method requires that we consider our proposal as a working hypothesis that should be questioned and subjected to further test. We have named the proposed process the ‘sugar oxidation cascade’ in recognition of the differing sugar compositions of the floral nectars ingested by hummingbirds and nectar bats in nature (Baker, 1975; Baker et al., 1998), as well as the lack of information concerning fructose metabolism in these animals. In humans, up to half of ingested fructose is converted by the liver to glucose, which then appears in the blood (Delarue et al., 1993). During exercise at 60% of VO2max, orally ingested glucose and fructose directly contribute up to 15 and 12%, respectively, to energy production (Adopo et al., 1994). In rat skeletal muscles, maximal rates of fructose transport are approximately eightfold lower than rates of glucose transport (Kristiansen et al., 1997). Upon entry into the sarcoplasm, fructose is phosphorylated and converted to glycogen, lactate or CO2, but these occur at rates far lower than when glucose is the precursor (Zierath et al., 1995). However, the predominant role of glucose in muscle energy metabolism in humans and rats should not lead to blind extrapolation to nectarivorous animals that ingest fructose – both as a monosaccharide and as part of the sucrose molecule – at much higher mass-specific rates than rats and humans.

An intriguing question concerns the role played by muscle glycogen. Upon entering muscle fibers and after phosphorylation by the hexokinase reaction, glucose can be converted to glycogen or broken down in the glycolytic pathway and oxidized. The flight muscles of hummingbirds and nectar bats are unique among vertebrate skeletal muscles in that they have hexokinase Vmax values so high that glucose oxidation can completely account for values during flight. Such high capacities for glucose phosphorylation are also observed in the flight muscles of frugivorous but not insectivorous bats (Yacoe et al., 1982). This indicates that sugar-rich diets and not just small body size or flight account for the high capacities for muscle glucose phosphorylation. By contrast, nectarivory is common to all hummingbird species; all hummingbird species we have examined so far display high hexokinase Vmax values in their flight muscles (M. J. Fernandez and R.S., unpublished). These activities are far higher than in other non-nectarivorous avian species (e.g. Bishop et al., 1995; Blomstrand et al., 1983; Crabtree and Newsholme, 1972) that have flight muscles with hexokinase Vmax values that are insufficient to allow glucose to serve as the sole oxidative fuel during flight. In possessing high enzymatic capacities as well as in displaying high physiological rates of glucose phosphorylation, hummingbird and nectar bat flight muscles appear rather similar to vertebrate hearts (Kashiwaya et al., 1994) that oxidize glucose and long-chain fatty acids in proportions that depend upon physiological circumstances (Collins-Nakai et al., 1994). Even while performing steady-state aerobic work, glycogen turnover occurs in hearts; glycogenolysis serves to buffer hexose phosphate concentrations and its contribution to the fueling of aerobic metabolism changes during transitions in work rate (Goodwin et al., 1995; Goodwin et al., 1998). Thus, when considering energy metabolism in the flight muscles of hummingbirds and nectar bats, there is good reason to doubt the validity of the assertion that “...the substrate of muscle glycolysis is glycogen, not glucose, and hexokinase is part of the glycogen synthesis pathway” (Fell, 2000), as well as to turn such doubt into further testable hypotheses.

Nectarivory and hovering flight were once traits found only among insects. In becoming hovering nectarivores, hummingbirds and nectar bats have converged, evolving enhanced capacities for O2 flux and sugar oxidation. Foraging behavior and metabolic regulation appear to have coevolved in rufous hummingbirds (Suarez and Gass, 2002; Suarez et al., 1990). Because foraging bouts tend to be brief in the wild (Diamond et al., 1986), such behavior would tend to ensure the oxidation of dietary sugar and to minimize the use of fat as a fuel. This avoids the inefficiency of a futile cycle involving the expenditure of energy to synthesize fat, followed by the breakdown of fat made from sugar to fuel further foraging. Instead, depletion of fat stores while foraging is avoided or minimized and oxidation of recently ingested sugar is favored.

Another advantage derived from the oxidation of sugar is the 15% higher yield of ATP per oxygen atom consumed (the P/O ratio) as compared with the oxidation of long-chain fatty acids (Brand, 2005). This suggests that, to support the energetic requirements of a unit mass of hummingbird during hovering, 15% lower is required when sugar is oxidized as compared with fat. This hypothesis is supported by results showing that declines by ∼15% as hummingbirds transition from the fasted state, when flight is fueled by fat, to the fed state, when flight is fueled by sugar (Welch et al., 2007). Many species of hummingbirds forage at relatively high altitude under conditions wherein flight performance might be adversely affected by hypobaric hypoxia (Altshuler and Dudley, 2002). Thus, we hypothesize that ingested sugar serves as a premium fuel for hummingbird flight.

Over the past half century, the field of comparative physiology has generated a fascinating and valuable body of knowledge concerning organismal function, physiological adaptation and patterns of functional variation across species. The exploration of functional biodiversity in the natural world continues to be central to the research agenda of the field. The research that led to the concept of the sugar oxidation cascade exemplifies how natural history, animal behavior and ecology can serve as the inspiration for comparative physiological questions. In turn, research in comparative physiology can lead to new concepts and greater understanding of the nature and extent of biodiversity. The answers to mechanistic, physiological questions continue to enrich the banquet at which ecologists, behaviorists and evolutionary biologists can feast.

The empirical work that led to this concept paper was supported by the US National Science Foundation (IOB 0517694) and UC MEXUS-CONACYT. For their support and collaboration at various stages in the history of these investigations, we are deeply grateful to P. W. Hochachka, J. R. B. Lighton, C. L. Gass, O. Mathieu-Costello, G. S. Brown, C. D. Moyes, B. H. Bakken and C. Martinez del Rio.

Adopo
E.
,
Peronnet
F.
,
Massicotte
D.
,
Brisson
G.
,
Hillaire-Marcel
C.
(
1994
).
Respective oxidation of exogenous glucose and fructose given in the same drink during exercise
.
J. Appl. Physiol.
76
,
1014
-
1019
.
Altshuler
D. L.
,
Dudley
R.
(
2002
).
The ecological and evolutionary interface of hummingbird flight physiology
.
J. Exp. Biol.
205
,
2325
-
2336
.
Baker
H.
(
1975
).
Sugar concentrations in nectars from hummingbird flowers
.
Biotropica
7
,
37
-
41
.
Baker
H.
,
Baker
I.
,
Hodges
S.
(
1998
).
Sugar composition of nectars and fruits consumed by birds and bats in the tropics and subtropics
.
Biotropica
30
,
559
-
586
.
Bartholomew
G. A.
,
Lighton
J. R. B.
(
1986
).
Oxygen consumption during hover-feeding in free-ranging Anna hummingbirds
.
J. Exp. Biol.
123
,
191
-
199
.
Beuchat
C. A.
,
Chaplin
S. B.
,
Morton
M. L.
(
1979
).
Ambient temperature and the daily energetics of two species of hummingbirds, Calypte anna and Selasphorus rufus
.
Physiol. Zool.
53
,
280
-
295
.
Bicudo
J. E. P. W.
,
Bianco
A. C.
,
Vianna
C. R.
(
2002
).
Adaptive thermogenesis in hummingbirds
.
J. Exp. Biol.
205
,
2267
-
2273
.
Bishop
C. M.
,
Butler
P. J.
,
Egginton
S.
,
El Haj
A. J.
,
Gabrielsen
G. W.
(
1995
).
Development of metabolic enzyme activity in locomotor and cardiac muscles of the migratory barnacle goose
.
Am. J. Physiol.
269
,
R64
-
R72
.
Blem
C. R.
(
1976
).
Patterns of lipid storage and utilization in birds
.
Am. Zool.
16
,
671
-
684
.
Blomstrand
E.
,
Challiss
R. A. J.
,
Cooney
G. J.
,
Newsholme
E. A.
(
1983
).
Maximal activities of hexokinase, 6-phosphofructokinase, oxoglutarate dehydrogenase, and carnitine palmitoyltransferase in rat and avian muscles
.
Biosci. Rep.
3
,
1149
-
1153
.
Brand
M. D.
(
2005
).
The efficiency and plasticity of mitochondrial energy transduction
.
Biochem. Soc. Trans.
33
,
897
-
904
.
Brooks
G. A.
(
1998
).
Mammalian fuel utilization during sustained exercise
.
Comp. Biochem. Physiol.
120B
,
89
-
107
.
Bursell
E.
,
Slack
E.
(
1976
).
Oxidation of proline by sarcosomes of the tsetse fly, Glossina morsitans
.
Insect Biochem.
6
,
159
-
167
.
Calder
W.
(
1987
).
Southbound through Colorado: migration of rufous hummingbirds
.
Natl. Geogr. Res.
3
,
40
-
51
.
Carpenter
F. L.
,
Paton
D. C.
,
Hixon
M. A.
(
1983
).
Weight gain and adjustment of feeding territory size in migrant hummingbirds
.
Proc. Natl. Acad. Sci. USA
80
,
7259
-
7263
.
Caviedes-Vidal
E.
,
Karasov
W. H.
,
Chediack
J. G.
,
Fasulo
V.
,
Cruz-Neto
A. P.
,
Otani
L.
(
2008
).
Paracellular absorption: a bat breaks the mammal paradigm
.
PloS ONE
3
,
e1425
.
Chai
P.
,
Millard
D.
(
1997
).
Flight and size constraints: hovering performance of large hummingbirds under maximal loading
.
J. Exp. Biol.
200
,
2757
-
2763
.
Chantler
P. D.
(
1982
).
Caged ATP set free in muscle
.
Nature
300
,
682
-
683
.
Collins-Nakai
R.
,
Noseworthy
D.
,
Lopaschuk
G.
(
1994
).
Epinephrine increases ATP production in hearts by preferentially increasing glucose metabolism
.
Am. J. Physiol.
267
,
H1862
-
H1871
.
Crabtree
B.
,
Newsholme
E. A.
(
1972
).
The activities of phosphorylase, hexokinase, phosphofructokinase, lactate dehydrogenase and the glycerol 3-phosphate dehydrogenases in muscles from vertebrates and invertebrates
.
Biochem. J.
126
,
49
-
58
.
Delarue
J.
,
Normand
S.
,
Pachiaudi
C.
,
Beylot
M.
,
Lamisse
F.
,
Riou
J. P.
(
1993
).
The contribution of naturally labelled 13C fructose to glucose appearance in humans
.
Diabetologia
36
,
338
-
345
.
di Prampero
P. E.
(
1985
).
Metabolic and circulatory limitations to VO2max at the whole animal level
.
J. Exp. Biol.
115
,
319
-
331
.
Diamond
J. M.
,
Karasov
W. H.
,
Phan
D.
,
Carpenter
F. L.
(
1986
).
Digestive physiology is a determinant of foraging bout frequency in hummingbirds
.
Nature
320
,
62
-
63
.
Dudley
R.
,
Winter
Y.
(
2002
).
Hovering flight mechanics of neotropical flower bats (Phyllostomidae: Glossophaginae) in normodense and hypodense gas mixtures
.
J. Exp. Biol.
205
,
3669
-
3677
.
Fell
D. A.
(
2000
).
Signal transduction and the control of expression of enzyme activity
.
Adv. Enzyme Regul.
40
,
35
-
46
.
Fleming
T. H.
,
Nunez
R. A.
,
Sternberg
D. S. L.
(
1993
).
Seasonal changes in the diets of migrant and non-migrant nectarivorous bats as revealed by carbon stable isotope analysis
.
Oecologia
94
,
72
-
75
.
Fons
R.
,
Sicart
R.
(
1976
).
Contribution a la connaissance du metabolisme energetique chez deux crocidurinae: Suncus etruscus (Savi, 1822) et Crocidura russula (Hermann, 1780) (Insectivora, Soricidae)
.
Mammalia
40
,
299
-
311
.
Fueger
P. T.
,
Bracy
D. P.
,
Malabanan
C. M.
,
Pencek
R. R.
,
Wasserman
D. H.
(
2004
).
Distributed control of glucose uptake by working muscles of conscious mice: roles of transport and phosphorylation
.
Am. J. Physiol.
286
,
E77
-
E84
.
Gass
C. L.
,
Romich
M. T.
,
Suarez
R. K.
(
1999
).
Energetics of hummingbird foraging at low ambient temperature
.
Can. J. Zool.
77
,
314
-
320
.
Goodwin
G. W.
,
Arteaga
J. R.
,
Taegtmeyer
H.
(
1995
).
Glycogen turnover in the isolated working rat heart
.
J. Biol. Chem.
270
,
9234
-
9240
.
Goodwin
G. W.
,
Taylor
C. S.
,
Taegtmeyer
H.
(
1998
).
Regulation of energy metabolism of the heart during acute increase in heart work
.
J. Biol. Chem.
273
,
29530
-
29539
.
Grinyer
I.
,
George
J.
(
1969
).
Some observations on ultrastructure of hummingbird pectoral muscles
.
Can. J. Zool.
47
,
771
-
774
.
Hammond
K. A.
,
Diamond
J. M.
(
1997
).
Maximal sustained energy budgets in humans and animals
.
Nature
386
,
457
-
462
.
Hermanson
J. W.
,
Ryan
J. M.
,
Cobb
M. A.
,
Bentley
J.
,
Schutt
W. A.
(
1998
).
Histochemical and electrophoretic analysis of the primary flight muscle of several phyllostomid bats
.
Can. J. Zool.
76
,
1983
-
1992
.
Hernandez
A.
,
Martinez del Rio
C.
(
1992
).
Intestinal disaccharidases in five species of phyllostomid bats
.
J. Comp. Physiol.
103B
,
105
-
111
.
Homsher
E.
(
1987
).
Muscle enthalpy production and its relationship to actomyosin ATPase
.
Annu. Rev. Physiol.
49
,
673
-
690
.
Homsher
E.
,
Kean
C.
(
1978
).
Skeletal muscle energetics and metabolism
.
Annu. Rev. Physiol.
40
,
93
-
131
.
Huang
S.
,
Czech
M. P.
(
2007
).
The GLUT4 glucose transporter
.
Cell Metab.
5
,
237
-
252
.
Jentjens
R. L. P. G.
,
Venables
M. C.
,
Jeukendrup
A. E.
(
2004
).
Oxidation of exogenous glucose, sucrose, and maltose during prolonged cycling exercise
.
J. Appl. Physiol.
96
,
1285
-
1291
.
Johansen
K.
(
1987
).
The world as a laboratory: physiological insights from Nature’s experiments
.
In Advances in Physiological Research
(ed.
McLennan
H.
,
Ledsome
J. R.
,
McIntosh
C. H. S.
), pp.
377
-
396
.
New York
:
Plenum Press
.
Jones
J. H.
(
1998
).
Optimization of the mammalian respiratory system: symmorphosis versus single species adaptation
.
Comp. Biochem. Physiol.
120B
,
125
-
138
.
Kashiwaya
Y.
,
Sato
K.
,
Tsuchiya
N.
,
Thomas
S.
,
Fell
D. A.
,
Veech
R. L.
,
Passonneau
J. V.
(
1994
).
Control of glucose utilization in working perfused rat heart
.
J. Biol. Chem.
269
,
25502
-
25514
.
Kristiansen
S.
,
Darakhshan
F.
,
Richter
E. A.
,
Handal
H. S.
(
1997
).
Fructose transport and GLUT5 protein in human sarcolemmal vesicles
.
Am. J. Physiol.
273
,
E543
-
E548
.
Lasiewski
R. C.
(
1963
).
Oxygen consumption of torpid, resting, active, and flying hummingbirds
.
Physiol. Zool.
36
,
122
-
140
.
Lopez-Calleja
M. V.
,
Bozinovic
F.
(
1995
).
Maximum metabolic rate, thermal insulation and aerobic scope in a small-sized Chilean hummingbird (Sephanoides sephanoides)
.
Auk
112
,
1034
-
1036
.
Lotz
C. N.
,
Martinez del Rio
C.
,
Nicholson
S. W.
(
2003
).
Hummingbirds pay a high cost for a warm drink
.
J. Comp. Physiol. Biochem. Syst. Environ. Physiol.
173
,
455
-
462
.
Maina
J. N.
(
2000
).
What it takes to fly: the structural and functional respiratory requirements in birds and bats
.
J. Exp. Biol.
203
,
3045
-
3064
.
Martinez del Rio
C.
(
1990
).
Sugar preferences in hummingbirds: the influence of subtle chemical differences on food choice
.
Condor
92
,
1022
-
1030
.
Mathieu-Costello
O.
,
Suarez
R. K.
,
Hochachka
P. W.
(
1992
).
Capillary-to-fiber geometry and mitochondrial density in hummingbird flight muscle
.
Respir. Physiol.
89
,
113
-
132
.
McNab
B. K.
(
1976
).
Seasonal fat reserves of bats in two tropical environments
.
Ecology
57
,
332
-
338
.
McNevin
D. B.
,
Badger
M. R.
,
Whitney
S. M.
,
von Caemmerer
S.
,
Tcherkez
G. B.
,
Farquhar
G. D.
(
2007
).
Differences in carbon isotope discrimination of three variants of D-ribulose-1,5-bisphosphate carboxylase/oxygenase reflect differences in their catalytic mechanisms
.
J. Biol. Chem.
282
,
36068
-
36076
.
McWhorter
T. J.
,
Martinez del Rio
C.
(
2000
).
Does gut function limit hummingbird food intake?
Physiol. Biochem. Zool.
73
,
313
-
324
.
McWhorter
T. J.
,
Bakken
B. H.
,
Karasov
W. H.
,
Martinez del Rio
C.
(
2006
).
Hummingbirds rely on both paracellular and carrier-mediated intestinal glucose absorption to fuel high metabolism
.
Biol. Lett.
2
,
131
-
134
.
Morales-Garza
M. R.
,
Del
C.
,
Arizmendi
M.
,
Campos
J. E.
,
Martinez-Garcia
M.
,
Valiente-Banuet
A.
(
2007
).
Evidences on the migratory movements of the nectar-feeding bat Leptonycterys curasoae in Mexico using random amplified polymorphic DNA (RAPD)
.
J. Arid Environ.
68
,
248
-
259
.
Newsholme
E. A.
,
Crabtree
B.
(
1986
).
Maximum catalytic activity of some key enzymes in provision of physiologically useful information about metabolic fluxes
.
J. Exp. Zool.
239
,
159
-
167
.
Odum
E. P.
,
Connell
C. E.
,
Stoddard
H. L.
(
1961
).
Flight energy and estimated flight ranges of some migratory birds
.
Auk
78
,
515
-
527
.
Pennycuick
C. J.
,
Rezende
M. A.
(
1984
).
The specific power output of aerobic muscle, related to the power density of mitochondria
.
J. Exp. Biol.
108
,
377
-
392
.
Peterson
C. C.
,
Nagy
K. A.
,
Diamond
J. M.
(
1990
).
Sustained metabolic scope
.
Proc. Natl. Acad. Sci. USA
87
,
2324
-
2328
.
Powers
D. R.
,
Nagy
K. A.
(
1988
).
Field metabolic rate and food consumption by free-living Anna’s hummingbirds (Calypte anna)
.
Physiol. Zool.
61
,
500
-
506
.
Rodriguez-Pena
N.
,
Stoner
K. E.
,
Schondube
J. E.
,
Ayala-Berdon
J.
,
Flores-Ortiz
C. M.
,
Martinez del Rio
C.
(
2007
).
Effects of sugar composition and concentration on food selection by Saussure’s long-nosed bat (Leptonycterys curasoae) and the long-tounged bat (Glossophaga soricina)
.
J. Mammal.
88
,
1466
-
1474
.
Schondube
J. E.
,
Herrera
M. L. G.
,
Martinez del Rio
C.
(
2001
).
Diet and the evolution of digestion and renal function in phyllostomid bats
.
Zoology
104
,
59
-
73
.
Seki
Y.
,
Sato
K.
,
Kono
T.
,
Abe
H.
,
Akiba
Y.
(
2003
).
Broiler chickens (Ross strain) lack insulin-responsive glucose transporter GLUT4 and have GLUT8 cDNA
.
Gen. Comp. Endocrinol.
133
,
80
-
87
.
Suarez
R. K.
(
1992
).
Hummingbird flight: sustaining the highest mass-specific metabolic rates among vertebrates
.
Experientia
48
,
565
-
570
.
Suarez
R. K.
(
1996
).
Upper limits to mass-specific metabolic rates
.
Annu. Rev. Physiol.
58
,
583
-
605
.
Suarez
R. K.
(
1998
).
Oxygen and the upper limits to animal design and performance
.
J. Exp. Biol.
201
,
1065
-
1072
.
Suarez
R. K.
,
Gass
C. L.
(
2002
).
Hummingbird foraging and the relation between bioenergetics and behaviour
.
Comp. Biochem. Physiol.
133A
,
335
-
343
.
Suarez
R. K.
,
Brown
G. S.
,
Hochachka
P. W.
(
1986
).
Metabolic sources of energy for hummingbird flight
.
Am. J. Physiol.
251
,
R537
-
R542
.
Suarez
R. K.
,
Lighton
J. R. B.
,
Moyes
C. D.
,
Brown
G. S.
,
Gass
C. L.
,
Hochachka
P. W.
(
1990
).
Fuel selection in rufous hummingbirds: ecological implications of metabolic biochemistry
.
Proc. Natl. Acad. Sci. USA
87
,
9207
-
9210
.
Suarez
R. K.
,
Lighton
J. R. B.
,
Brown
G. S.
,
Mathieu-Costello
O.
(
1991
).
Mitochondrial respiration in hummingbird flight muscles
.
Proc. Natl. Acad. Sci. USA
88
,
4870
-
4873
.
Suarez
R. K.
,
Staples
J. F.
,
Lighton
J. R. B.
,
West
T. G.
(
1997
).
Relationships between enzymatic flux capacities and metabolic flux rates in muscles: nonequilibrium reactions in muscle glycolysis
.
Proc. Natl. Acad. Sci. USA
94
,
7065
-
7069
.
Suarez
R. K.
,
Welch
K. C.
Jr
,
Hanna
S. K.
,
Herrera
M. L. G.
(
2009
).
Flight muscle enzymes and metabolic flux rates during hovering flight of the nectar bat, Glossophaga soricina: further evidence of convergence with hummingbirds
.
Comp. Biochem. Physiol.
153A
,
136
-
140
.
Sweazea
K. L.
,
Braun
E. J.
(
2006
).
Glucose transporter expression in English sparrows (Passer domesticus)
.
Comp. Biochem. Physiol.
144B
,
263
-
270
.
Szentesi
P.
,
Zaremba
R.
,
van Mechelen
W.
,
Stienen
G. J. M.
(
2001
).
ATP utilization for calcium uptake and force production in different types of human skeletal muscle fibers
.
J. Physiol.
531
,
393
-
403
.
Taylor
C. R.
(
1987
).
Structural and functional limits to oxidative metabolism: insights from scaling
.
Annu. Rev. Physiol.
49
,
135
-
146
.
Voigt
C. C.
,
Speakman
J. R.
(
2007
).
Nectar-feeding bats fuel their high metabolism directly with exogenous carbohydrates
.
Funct. Ecol.
21
,
913
-
921
.
Voigt
C. C.
,
Winter
Y.
(
1999
).
Energetic cost of hovering flight in nectar-feeding bats (Phyllostomidae: Glossophaginae) and its scaling in moths, birds and bats
.
J. Comp. Physiol.
169
,
38
-
48
.
Wagner
P. D.
(
1996
).
A theoretical analysis of factors determining VO2max at sea level and altitude
.
Respir. Physiol.
106
,
329
-
343
.
Weber
J.-M.
,
Haman
F.
(
2004
).
Oxidative fuel selection: adjusting mix and flux to stay alive
.
Int. Congr. Ser.
1275
,
22
-
31
.
Weber
J.-M.
,
Roberts
T. J.
,
Vock
R.
,
Weibel
E. R.
,
Taylor
C. R.
(
1996
).
Design of the oxygen and substrate pathways. III. Partitioning energy provision from carbohydrates
.
J. Exp. Biol.
199
,
1659
-
1666
.
Weibel
E. R.
(
1984
).
The Pathway for Oxygen
.
Cambridge, MA
:
Harvard University Press
.
Weibel
E. R.
,
Taylor
C. R.
,
Gehr
P.
,
Hoppeler
H.
,
Mathieu
O.
,
Maloiy
G. M. O.
(
1981
).
Design of the mammalian respiratory system. IX. Functional and structural limits for oxygen flow
.
Respir. Physiol.
44
,
151
-
164
.
Weibel
E. R.
,
Taylor
C. R.
,
Weber
J.-M.
,
Vock
R.
,
Roberts
T. J.
,
Hoppeler
H.
(
1996
).
Design of the oxygen and substrate pathways. VII. Different structural limits for oxygen and substrate supply to muscle mitochondria
.
J. Exp. Biol.
199
,
1699
-
1709
.
Welch
K. C.
,
Suarez
R. K.
(
2007
).
Oxidation rate and turnover of ingested sugar in hovering Anna’s (Calypte anna) and rufous (Selasphorus rufus) hummingbirds
.
J. Exp. Biol.
210
,
2154
-
2162
.
Welch
K. C.
,
Suarez
R. K.
(
2008
).
Altitude and temperature effects on the energetic cost of hover-feeding in migratory rufous hummingbirds, Selasphorus rufus
.
Can. J. Zool.
86
,
161
-
169
.
Welch
K. C.
,
Bakken
B. H.
,
Martinez del Rio
C.
,
Suarez
R. K.
(
2006
).
Hummingbirds fuel hovering flight with newly ingested sugar
.
Physiol. Biochem. Zool.
79
,
1082
-
1087
.
Welch
K. C.
,
Altschuler
D. L.
,
Suarez
R. K.
(
2007
).
Oxygen consumption rates in hovering hummingbirds reflect substrate-dependent differences in P/O ratios: carbohydrate as a ‘premium fuel’
.
J. Exp. Biol.
210
,
2146
-
2153
.
Welch
K. C.
,
Herrera
L. G.
,
Suarez
R. K.
(
2008
).
Dietary sugar as a direct fuel for flight in the nectarivorous bat, Glossophaga soricina
.
J. Exp. Biol.
211
,
310
-
316
.
Wells
D.
(
1993
).
Muscle performance in hovering hummingbirds
.
J. Exp. Biol.
178
,
39
-
57
.
Winter
Y.
,
Voigt
C.
,
von Helversen
O.
(
1998
).
Gas exchange during hovering flight in a nectar-feeding bat Glossophaga soricina
.
J. Exp. Biol.
201
,
237
-
244
.
Wisniewski
E.
,
Gellerich
F. N.
,
Kunz
W. S.
(
1995
).
Distribution of flux control among the enzymes of mitochondrial oxidative phosphorylation in calcium-activated saponin-skinned rat musculus soleus fibers
.
Eur. J. Biochem.
230
,
549
-
554
.
Yacoe
M. E.
,
Cummings
J. W.
,
Myers
P.
,
Creighton
G. K.
(
1982
).
Muscle enzyme profile, diet and flight in South American bats
.
Am. J. Physiol.
242
,
R189
-
R194
.
Zierath
J. R.
,
Nolte
L. A.
,
Wahlstrom
E.
,
Galuska
D.
,
Shepherd
P. R.
,
Kahn
B. B.
,
Wallberg-Henriksson
H.
(
1995
).
Carrier-mediated fructose uptake significantly contributes to carbohydrate metabolism in human skeletal muscle
.
Biochem. J.
311
,
517
-
521
.