You are on page 1of 7

Am J Physiol Regul Integr Comp Physiol 289: R895–R901, 2005;

doi:10.1152/ajpregu.00641.2004. Letters to the Editor

The following is the abstract of the article discussed in the consumption of protons by the Lohmann reaction (17); this
subsequent letter: term refers to the hydrolysis of ATP and its simultaneous
regeneration at the expense of phosphocreatine, catalyzed by
Robergs, Robert A., Farzenah Ghiasvand, and Daryl
Parker. Biochemistry of exercise-induced metabolic acidosis. creatine kinase (7, 17, 21) and is stoichiometrically equivalent
Am J Physiol Regul Integr Comp Physiol 287: R502–R516, 2004 to the splitting, or hydrolysis, of phosphocreatine to inorganic
doi:10.1152/ajpregu.00641.2004.—The development of acidosis phosphate [all these terms can be found in the literature, but
during intense exercise has traditionally been explained by the in- note that this reaction does not occur (21), being only a name
creased production of lactic acid, causing the release of a proton and for the sum of two enzyme-catalyzed processes]. This proton
the formation of the acid salt sodium lactate. On the basis of this consumption is sometimes called dynamic (23) or metabolic
explanation, if the rate of lactate production is high enough, the (21) buffering. Taking the fall in cytosolic pH (to use a
cellular proton buffering capacity can be exceeded, resulting in a physical analogy) as the strain resulting from the glycolytic
decrease in cellular pH. These biochemical events have been termed proton-load stress, the ratio of lactate increase to pH fall has
lactic acidosis. The lactic acidosis of exercise has been a classic been used as an apparent buffer capacity, a measure of both
explanation of the biochemistry of acidosis for more than 80 years.
This belief has led to the interpretation that lactate production causes
these components of proton handling (23) More properly, the
acidosis and, in turn, that increased lactate production is one of the true physicochemical buffer capacity is the ratio to the pH fall
several causes of muscle fatigue during intense exercise. This review of the net proton load, that is, the total glycogenolytic proton
presents clear evidence that there is no biochemical support for lactate load less that consumed by the Lohmann reaction (16).
production causing acidosis. Lactate production retards, not causes, In criticizing this approach, Robergs et al. (21) note that the
acidosis. Similarly, there is a wealth of research evidence to show that glycogenolytic proton load is generated by the hydrolysis of
acidosis is caused by reactions other than lactate production. Every ATP rather than ionization of lactic acid (6, 12, 14), while the
time ATP is broken down to ADP and Pi, a proton is released. When reduction of pyruvate to lactate actually consumes protons,
the ATP demand of muscle contraction is met by mitochondrial mitigating acidification rather than causing it. Thus the naı̈ve
respiration, there is no proton accumulation in the cell, as protons are notion that glycolysis adds lactic acid to the cell is untenable
used by the mitochondria for oxidative phosphorylation and to main- (21). Robergs et al. develop this criticism to argue for a
tain the proton gradient in the intermembranous space. It is only when
the exercise intensity increases beyond steady state that there is a need
different understanding of acid-base balance, one implication
for greater reliance on ATP regeneration from glycolysis and the of which is that for biopsy and 31P MRS studies of muscle
phosphagen system. The ATP that is supplied from these nonmito- energetics, the traditional approach yields gross underestima-
chondrial sources and is eventually used to fuel muscle contraction tions of the true muscle buffer capacity (21).
increases proton release and causes the acidosis of intense exercise. I will argue that while this analysis of proton generation is
Lactate production increases under these cellular conditions to prevent broadly correct at resting pH, it is much less so at low pH
pyruvate accumulation and supply the NAD⫹ needed for phase 2 of values commonly found in exercising muscle. Furthermore,
glycolysis. Thus increased lactate production coincides with cellular because this analysis of muscle cell buffering capacity adds
acidosis and remains a good indirect marker for cell metabolic together components of proton generation that are partially
conditions that induce metabolic acidosis. If muscle did not produce cancelled by processes of proton consumption, it may be
lactate, acidosis and muscle fatigue would occur more quickly and
physiologically misleading.
exercise performance would be severely impaired.
BIOCHEMICAL BACKGROUND
Lactate accumulation, proton buffering, and pH change in
ischemically exercising muscle To understand this, we must consider the stoichiometry (Table
1) and then examine its implications for the analysis of cellular
By reviewing theory (6, 11, 12, 14) and reanalyzing data (3, 15, buffering (Table 2). In ischemic exercise ATP is produced by
19, 24, 25), Robergs and colleagues argue (21), among other glycogenolysis to lactate and by the Lohmann reaction. The
things, for a particular approach to the generation and disposal effects on cytosolic pH (which can vary from ⬃7 at rest to ⬃6
of protons in exercising skeletal muscle. They rightly insist that in intense exercise) depend on the pH-dependent proton stoi-
interpreting muscle cell pH changes requires an appropriate chiometries of the component processes (1, 2, 6, 7, 12, 14, 16,
analysis of proton stoichiometry. I would add two points. First, 17). Table 1 shows (centered) the biochemical equations in a
one may be misled by neglecting, even for illustrative pur- form that takes account of this pH dependence and cites the
poses, the pH dependence of this stoichiometry. Second, care- corresponding expressions or figures in Robergs et al. (21)
ful accounting is needed to assess how physicochemical buff- where, for a simpler exposition, this complication is ignored. (I
ering contributes to cellular acid-base balance. To argue this, I follow Robergs et al. in not using the alternative strong ion
will consider only ischemic exercise (exercise without blood difference formalism, which is not very useful when lactate is
flow), which has the simplifying advantages that mitochondrial the only strong ion measured.) Table 1 also shows algebraic
proton uptake and cellular proton efflux (21) are negligible, expressions for the proton stoichiometries, along with the
and the only glycolytic substrate is glycogen (although for numerical values stated or implicit in Robergs et al. (21) and as
completeness, results for glucose will also be given). calculated from the present analysis for three pH values span-
What might be called the traditional analysis criticized by ning the physiological range. Table 2 gives expressions for the
Robergs et al. takes the increase in cytosolic lactate concen- estimation of various versions of calculated buffer capacity.
tration as a measure of the proton load arising from glycogen- Figure 1 illustrates the application to a data set (16).
olysis, which is disposed of in two ways (14, 16). The first is Glycolysis from glycogen to pyruvate (Table 2 in Ref. 21)
physicochemical buffering [also called static (23) or structural generates about one proton per pyruvate at resting pH, increas-
(21) buffering], predominantly by cytosolic imidazole groups ing as pH falls (Eq. 1b); this is not very different for glycolysis
and inorganic phosphate (1, 2, 16, 23, 26). The second is the from glucose (although in the simplified Robergs analysis the
http://www.ajpregu.org 0363-6119/05 $8.00 Copyright © 2005 the American Physiological Society R895
Downloaded from journals.physiology.org/journal/ajpregu (210.050.069.126) on June 26, 2022.
Letters to the Editor
R896
Table 1. Components of proton (H⫹) generation and consumption in ischaemically exercising muscle
Stoichiometry in Robergs et al. (21) and the Present Analysis†

Process and Analysis* Component Substrate Robergs pH 7 pH 6.4 pH 6 Equation and Derivation

Glycolysis to pyruvate: Eqs. 1 and 2 and Table 2 in Robergs et al. (21)


X ⫹ 2␻ADPa ⫹ 2␻Pib ⫹ 2NAD⫹ 3 Y ⫹ 2Pyr⫺ ⫹ 2␻ATPc ⫹ 2NADH ⫹ 2␻H2O ⫹ 2[␻(a ⫹ b ⫺ c) ⫹ 2]H⫹
ATP production ⫽ ␻ per pyruvate ⫽ ␻⌬La Eq. 1a
H⫹ production ⫽ [␻(a ⫹ b ⫹ c) ⫹ 2]⌬La H⫹ produced per lactate‡ glycogen (␻ ⫽ 3/2) 0.5 0.91 1.42 1.93 Eq. 1b
glucose (␻ ⫽ 1) 1 1.27 1.62 1.95

Reduction of pyruvate to lactate Fig. 9 in Robergs et al. (21)


Pyr⫺ ⫹ NADH ⫹ H⫹ 3 La⫺ ⫹ NAD⫹
H⫹ consumption ⫽ ⌬La H⫹ consumed per lactate glycogen and glucose 1 1 1 1 Eq. 2

Glycolysis to lactate (i.e., glycolysis to pyruvate) and reduction of pyruvate to lactate: Eqs. 3 and 4 in Robergs et al. (21)
X ⫹ 2␻ADPa ⫹ 2␻Pib 3 Y ⫹ 2La⫺ ⫹ 2␻ATPc ⫹ 2␻H2O ⫹ 2[␻(a ⫹ b ⫺ c) ⫹ 1]H⫹
H⫹ production ⫽ [␻(a ⫹ b ⫺ c) ⫹ 1]⌬La H⫹ produced per lactate glycogen ⫺0.5 ⫺0.09 0.42 0.93 Eq. 3 (from Eqs. 1b and 2)
glucose 0 0.27 0.62 0.95

Hydrolysis of ATP: Fig. 10 and Eq. 7 in Robergs et al. (21)


ATPc ⫹ H2O 3 ADPa ⫹ Pib ⫹ (c ⫺ a ⫺ b)H⫹
H⫹ production: in general ⫽ (c ⫺ a ⫺ b) H⫹ produced per ATP glycogen and glucose 1 0.73 0.38 0.05 Eq. 4a
per ATP
from glycolytic ATP ⫽ ␻(c ⫺ a ⫺ b)⌬La H⫹ produced per lactate glycogen 1.5 1.09 0.58 0.07 Eq. 4b (from Eqs. 1a and 4a)
glucose 1 0.73 0.38 0.05
from ATP via CK ⫽ (c ⫺ a ⫺ b)⌬PCr H⫹ produced per PCr glycogen and glucose 1 0.73 0.38 0.05 Eq. 4c (from Eqs. 4a and 6a)

Glycolysis to lactate coupled with ATP hydrolysis: Eqs. 5 and 6 in Robergs et al. (21)
X⫺ 3 Y ⫹ 2La⫺ ⫹ 2H⫹
H⫹ production ⫽ ⌬La H⫹ produced per lactate glycogen and glucose 1 1 1 1 Eq. 5a (from Eqs. 3 and 4b)

Fraction of this H derived from ATP glycogen 1.5 1.09 0.58 0.07 0.05 Eq. 5b (from Eqs. 4b and 5a)
hydrolysis ⫽ ␻(c ⫺ a ⫺ b)
glucose 1 0.73 0.38 0.05

ATP regeneration by phosphocreatine, catalysed by CK: Fig. 4 in Robergs et al. (21)


PCrg ⫹ ADPa ⫹ (c ⫺ a ⫺ g)H⫹ 3 Cr ⫹ ATPc
ATP production ⫽ 1 ATP per PCr Eq. 6a
H⫹ consumption ⫽ (c ⫺ a ⫺ g)⌬PCr H⫹ consumed per ATP glycogen and glucose 1 0.93 0.82 0.70 Eq. 6b

Lohmann reaction: ATP hydrolysis plus ATP regeneration: Figs. 4 and 11 in Robergs et al. (21)
PCrg ⫹ (b ⫺ g)H⫹ ⫹ H2O 3 Pib ⫹ Cr
H⫹ consumption ⫽ (b ⫺ g)⌬PCr H⫹ consumed per ATP glycogen and glucose 0 0.20 0.44 0.65 Eq. 7 (from Eqs. 4c and 6b)
Pyr, pyruvate; La, lactate; PCr, phosphocreatine; Cr, creatine; ADP, adenosine diphosphate. In the biochemical equations (centered) ATP, ADP, PCr, and Pi
refer to the sum of all relevant chemical species, and the superscripts a, b, c, g are the generally nonintegral average charges of ADP, Pi, ATP and PCr (Cr is
uncharged), respectively. In the algebraic equations, concentration brackets are omitted for simplicity; ⌬La denotes an increase and ⌬PCr and ⌬pH a decrease
during exercise, and the stoichiometric coefficients a, b, c, g are simply the charges in the biochemical equations. The proton stoichiometry of these equations
follows from the electroneutrality condition. The conclusions here follow directly from this, relatively independent of the numerical values of the charges. The
equations apply to both glycolytic substrates: for glycogen (the predominant substrate in ischemic exercise), the symbols X and Y refer to glycogenn and
glycogenn⫺1 and the ATP stoichiometric factor ␻ ⫽ 3/2; when X refers to glucose, Y is nothing and ␻ ⫽ 1. ␤ is the physicochemical cytosolic buffer capacity.
Eqs. 1–7 in Table 1 are based mainly on Refs. 12, 14, 17, Eqs. 8 –12 in Table 2 mainly on Refs. 1, 23, 26; see Ref. 16 for more details. Equations 13–15 in
Table 2 from Ref. 21 are translated into the present terminology using the analysis presented here. One complication ignored in these expressions, although not
in the illustrative calculations in Fig. 1, is that the pH dependence of both stoichiometric coefficients and buffer capacity requires that, for example, (b ⫺ g)⌬PCr
and ␤⌬pH should be (b ⫺ g)dPCr and ␤dpH, or in practice their finite-sum equivalents, ⌺[(b ⫺ g)␦PCr] and ⌺[␤␦pH]. *Reference is made to the numbered
equations and figures in Robergs et al. (21), where for simplicity, pH effects on stoichiometry are ignored, which are explicit here. Equations 1– 4 (21) ignore
the charges on ADP, ATP, Pi, lactate, and pyruvate, and Figs 4, 10, and 11 in (21) show only ADP3⫺, ATP4⫺, Pi2⫺ and PCr2⫺ (see footnote†). †These
stoichiometric coefficients are calculated using the analysis of Kushmerick (17), in which ␣, ⫺␤, and ⫺␥ are equivalent to c ⫺ a ⫺ b, c ⫺ a ⫺ g and b ⫺ g
here (this ␤ is not that used here for buffer capacity); these coefficients can be calculated empirically from pH using ␥ ⫽ 27.239 ⫺ 13.593pH ⫹ 2.1440(pH)2 ⫺
0.10887(pH)3, ␤ ⫽ ⫺11.869 ⫹ 5.6685pH ⫺ 0.92213(pH)2 ⫹ 0.047950(pH)3 and ␣ ⫽ 39.108 ⫺ 19.262pH ⫹ 3.0662(pH)2 ⫺ 0.15682(pH)3 (17). Values are
given for pH 7 (rest), pH 6.5 (end of exercise in Figure 1) and pH 6 (an extreme pH which can be reached in intense exercise); they are near-linear with pH
over this range. Values of the coefficients implicit in Robergs’ analysis are calculated by setting a ⫽ ⫺3, b ⫽ g ⫽ ⫺2, c ⫽ ⫺4, so that c ⫺ a ⫺ b ⫽ c ⫺ a ⫺
g ⫽ 1 and b ⫺ g ⫽ 0 (see footnote*). Glycogen and glucose substrate are distinguished where necessary. The conclusions presented here are not very sensitive
to the exact values chosen for these coefficients. ‡In Eq. 1 the product is pyruvate, but the expression is given for protons per lactate, because what is usually
measured (24, 25) or inferred (16) is accumulation of lactate, which in ischemic exercise is very close to lactate ⫹ pyruvate.

AJP-Regul Integr Comp Physiol • VOL 289 • SEPTEMBER 2005 • www.ajpregu.org


Downloaded from journals.physiology.org/journal/ajpregu (210.050.069.126) on June 26, 2022.
Letters to the Editor
R897
Table 2. Analysis of H⫹ buffering in ischaemically exercising muscle
Processes and Equations Equation and Derivation

Physicochemical version of traditional analysis


Net H⫹ load to cytosolic buffers ⫽ ⌬[La] ⫺ (b ⫺ g)⌬PCr ⫽ (glycolytic H⫹ load) ⫺ (Lohmann-consumed H⫹)
In the absence of glycolysis*; net H⫹ load ⫽ (b ⫺ g)⌬PCr, which is negative Eq. 8 (from Eqs. 5a and 7)
H⫹ buffered ⫽ ␤⌬pH Eq. 9
f Estimated physicochemical buffer capacity* ⫽ (net H⫹ load)/⌬pH ⫽ ␤ Eq. 10 (from Eq. 9)
Metabolic version of traditional analysis†
Metabolic H⫹ load ⫽ H⫹ production by glycolysis coupled with ATP hydrolysis ⫽ ⌬La Eq. 5a
H⫹ consumption ⫽ buffered ⫹ consumed by ATP regeneration from phosphocreatine ⫽ ␤⌬pH ⫹ (b ⫺ g)⌬PCr Eq. 11 (from Eqs. 7 and 9)
f Apparent buffer capacity ⫽ (glycolytic H⫹ load)/(pH fall)‡ ⫽ ⌬[La]/⌬pH Eq. 12a (from Eqs. 5a and 11)
f Ratio to true ␤ ⫽ 1 ⫹ (b ⫺ g)⌬PCr/(␤⌬pH) ⫽ 1 ⫹ (Lohmann-consumed H⫹)/(buffered H⫹) Eq. 12b (from Eqs. 9 and 12a)
Total muscle buffering analysis‡ of Robergs et al. (21)
H⫹ production ⫽ H⫹ from ATP hydrolysis ⫹ H⫹ from pyruvate synthesis ⫽ (c ⫺ a ⫺ b)[⌬PCr ⫹ ␻⌬La] ⫹
[␻(a ⫹ b ⫺ c) ⫹ 2]⌬La ⫽ 2⌬La ⫹ (c ⫺ a ⫺ b)⌬PCr Eq. 13 (from Eqs. 1b, 4b, and 4c)
H⫹ consumption ⫽ H⫹ buffered ⫹ H⫹ consumed by lactate reduction ⫹ H⫹ consumed by phosphocreatine
breakdown ⫽ ␤⌬pH ⫹ ⌬La ⫹ (c ⫺ a ⫺ g)⌬PCr Eq. 14 (from Eqs. 2, 6b, and 9)
f Proposed total buffer capacity ⫽ (total H⫹ production)/(pH fall) Eq. 15a (from Eqs. 9 and 13)
f Ratio to ␤ ⫽ 2 ⫹ [(b ⫹ c ⫺ a ⫺ 2g)⌬PCr/(␤⌬pH)] ⫽ 2 ⫹ [1 ⫹ (c ⫺ a ⫺ g)/(b ⫺ g)](Lohmann-consumed
H⫹)/(buffered H⫹) Eq. 15b (from Eqs. 9 and 15a)
See Table 1 for terminology, abbreviations, and biochemical background, and for the literature sources of the equations in this table; see the body of Table
1 for Eqs. 1– 7. *For example, in very early exercise (1, 4; see Fig 1C). †More complex expressions can take account of changes in other metabolites at lower
concentrations than lactate (13, 23, 25). ‡The relationship between these estimates of buffer capacity depends on the ratio of Lohmann-consumed to buffered
protons (Eq. 12b). This is ⫺1, when glycolysis is negligible, for example, in very early exercise where pH rises (see Fig 1C). It is undefined for pH ⬇ 0, then
achieves a value in later exercise that depends on metabolic regulation (see text), reaching about 0.4 in Fig. 1 (16), at which point the apparent buffer capacity
of Eq. 12a (23) is 40% higher than true physicochemical buffer capacity (Fig 1B), while the total buffer capacity of Eq. 15a (21) is about 3 times higher (Fig
1A).

assumed stoichiometric coefficients differ by a factor of 2). coupled ATP hydrolysis (Eq. 5b) is therefore numerically
Next, reduction of pyruvate to lactate consumes one proton, equal to the proton load from hydrolysis of glycolytic ATP
independent of pH (Eq. 2). At resting pH, because this disposes (Eq. 4b): only in the Robergs analysis for glucose is this
of protons roughly equivalent to the protons generated earlier fraction unity (21) (Table 1).
in glycolysis, both glycogenolysis and glycolysis to lactate (Eq. Now, we must introduce creatine kinase. In muscle, over a
3) generate few protons (21) [in the simplified Robergs anal- wide range of ATP turnover, the regeneration of ATP at the
ysis, glycolysis from glucose to lactate produces none, while expense of phosphocreatine (Eq. 6a) buffers ATP against
glycogenolysis consumes one proton per glucosyl (21)]. The temporary imbalance between ATP use and glycogenolytic
ATP generated by glycolysis does not accumulate, being si- ATP production (9). This, the Lohmann reaction (see previous
multaneously hydrolyzed by ATPases, resulting in the gener- section), consumes a number of protons equal to the difference
ation of almost one proton per ATP at resting pH, decreasing in charge between phosphocreatine and inorganic phosphate
as pH falls (Eq. 4a). Although two of the products of ATP (Eq. 7). On the simplest analysis in which the phosphocreatine
hydrolysis, inorganic phosphate and ADP, are recycled by has charge ⫺2, this is just the monoanion fraction of inorganic
glycolysis, the protons accumulate to acidify the cell (12, 21). phosphate, 1/[1⫹10(pH⫺6.8)] ⬇ 0.4 at resting pH (1, 14, 16, 26),
The result is that glycogenolysis or glycolysis to lactate cou- but more recent analysis, taking account of hitherto unconsid-
pled to simultaneous ATP hydrolysis generates exactly one ered potassium binding, suggests a smaller value of ⬃0.2 at
proton per lactate, independent of pH (Eq. 5a) (12). rest, increasing as pH falls (7, 17). There is some disagreement
It is the main premise of Robergs et al. (21) that the principal (16) about which value should be used (compare e.g., Refs. 1,
source of the proton arising from coupled glycolysis is ATP 4, 10, 20, and 22), but this is not important to the present
hydrolysis rather than lactic acid synthesis per se. Table 1 argument. However, it is a feature of the illustrative stoichi-
shows that this is true at resting pH (Eq. 5b); however, as pH ometry in the Robergs analysis (21), where this whole ap-
falls, this proportion is reversed as proton production by ATP proach is rejected, that the implied value of the Lohmann
hydrolysis decreases (6, 12, 14); until pH 6, most of the coefficient is zero (see Table 1, †footnote).
glycolytic protons actually do come from glycolysis to lactate
(Eq. 5b). Thus this claim (21) is a significant oversimplification “ TRADITIONAL” APPROACHES TO INTRACELLULAR
PROTON BUFFERING
as a result of neglecting the correct values of the stoichiometric
coefficients. Note that this argument is quite robust. To calcu- This is summarized in Table 2. A standard interpretation (7, 14,
late the protons produced by glycogenolysis coupled to ATP 16, 23, 25, 26) is that glycogenolysis generates protons, which
hydrolysis (Eq. 5a), we simply add the protons produced by are buffered by passive processes and consumed by the Loh-
glycogenolysis to lactate (Eq. 3) to those produced by the mann reaction (Eq. 7). What the Lohmann reaction leaves
balancing hydrolysis of the ATP so formed (Eq. 4b); the behind is the net proton load (Eq. 8) that produces a pH change
pH-dependent proton stoichiometry cancels out, so the result is depending on the physicochemical buffer capacity (Eq. 9) (16);
one proton per lactate at all pH values, independent of sub- this can, therefore, be estimated as their ratio (Eq. 10), for
strate, and also using the coefficients assumed by Robergs et al. comparison with ex vivo measurements (1, 5, 16, 17). Follow-
(21) (Table 1). The fraction of glycolytic protons supplied by ing Sahlin’s terminology for the components of buffering (23),
AJP-Regul Integr Comp Physiol • VOL 289 • SEPTEMBER 2005 • www.ajpregu.org
Downloaded from journals.physiology.org/journal/ajpregu (210.050.069.126) on June 26, 2022.
Letters to the Editor
R898
I call this the physicochemical approach. The underlying con- capacity (23, 25), which exceeds the true physicochemical
cept is of a net proton-generator (or, in early exercise, proton- buffer capacity (Eq. 12b) (there is some ambiguity about this
consumer) comprising glycolysis, ATPase and creatine kinase relationship in the older experimental literature). Following
added to a cytosol containing only passive buffers. The argu- Robergs’ terminology (21) for the components of buffering, I
ment is spelled out by Kemp and colleagues (16). call this the metabolic approach. The underlying concept is of
An alternative version of this approach is to regard glyco- a net proton-generator comprising glycolysis plus ATPase
genolysis as a source of protons (Eq. 5a) that are dealt with added to a cytosol containing passive buffers and the net
both by physicochemical buffering and metabolically by the proton-consuming creatine kinase system. Thus the conceptual
Lohmann reaction (Eq. 11). The ratio of the glycogenolytic difference between the apparent and true buffer capacities is
proton load (i.e., the lactate increase) to the pH fall (Eq. 12a) where the boundary is drawn between the generation and
can conveniently be taken as a functional or apparent buffer disposal of protons, whether creatine kinase is lumped with the

AJP-Regul Integr Comp Physiol • VOL 289 • SEPTEMBER 2005 • www.ajpregu.org


Downloaded from journals.physiology.org/journal/ajpregu (210.050.069.126) on June 26, 2022.
Letters to the Editor
R899
ATP-generating enzymes of glycolysis or with the proton- applied to published data from exercising human quadriceps
buffers of the cytosol. (3, 19, 24, 25). In particular, they argue (in their discussion of
The importance of these distinctions is that the balance their Figs. 16 and 17) that 1) given the known cytosolic buffer
between proton consumption and proton buffering depends on capacity, the proton balance equations make sense only for the
metabolic regulation (23), specifically, the relationship be- new model; and 2) the true buffer capacity in vivo is much
tween glycogen phosphorylase flux and the concentration of its greater than the traditional view allows (21). However, com-
substrate, phosphate (16). Because of the stoichiometric parison of estimates of buffer capacity inferred in vivo with
Lohmann-reaction relationship between phosphocreatine and those obtained by titration of muscle homogenate in vitro is by
phosphate, this can be seen as a negative feedback mechanism, no means straightforward (1, 16, 17), and cannot therefore
whereby phosphate is an error signal responding to the time- settle this question. Insufficient detail is given of these calcu-
integrated mismatch between ATP usage and glycogenolytic lations (21) to test them directly against the analysis above (see
ATP production (16). Although its role in the regulation of
Appendix). However, the main issue is more fundamental, and
glycogenolysis is controversial (7, 8, 18), this phosphate de-
can be argued on theoretical grounds.
pendence of phosphorylase is important in the control of pH,
and therefore in the relationship (Eq. 12b) between apparent Robergs et al. suggest that total proton generation (Eq. 13)
(Eq. 12a) and true buffer capacity (Eq. 10). One reason for the has two components (Fig. 1A). The first is proton generation by
early popularity (23) of the apparent buffer capacity was that ATP hydrolysis (Eq. 4a), in which one can distinguish the
the system operates so that the relationship between pH and protons generated by hydrolysis of ATP produced from phos-
lactate is remarkably linear and invariant [see Fig. 3 in (21)]: phocreatine via creatine kinase (Eq. 4c) from those generated
the constraints on this, and its implications for metabolic from ATP produced by glycogenolysis (Eq. 4b). The second
control, are discussed elsewhere (16). main component of proton generation (21) is glycogenolysis to
pyruvate (Eq. 1b). Total proton generation is independent of
PROPOSED NEW APPROACH TO INTRACELLULAR
PROTON BUFFERING substrate (Eq. 13). Robergs et al. (21) divide total proton
consumption (Eq. 14) into three components. The first (Fig.
This is also summarized in Table 2. Robergs et al. (21) argue 1A) is the traditional passively buffered component (Eq. 9).
that the traditional analysis is wrong because the protons The second represents the buffering of glycolytic protons by
produced during glycogenolysis arise from ATP hydrolysis lactate formation (Eq. 2). The third represents protons con-
(Eq. 4a) not lactic acid. However, we have seen that the sumed in phosphocreatine breakdown; as the acid-base effects
premise is only approximately true and progressively less so as
of ATP hydrolysis matching phosphocreatine breakdown are
pH falls (Eq. 5b). Furthermore, the traditional analysis makes
already taken into account in Eq. 9, this must be the unidirec-
no necessary assumption about the ultimate source of the
protons in Eq. 5a; references in the literature to protons arising tional reaction (Eq. 6b) rather than the net Lohmann reaction
from lactic acid can safely be replaced by the more conceptu- (Eq. 7). Robergs et al. argue that the correct, or at least most
ally accurate protons accompanying lactate. physiologically meaningful, calculation of buffer capacity is
More fundamentally, Robergs et al. argue for a new analysis the ratio of total proton production to pH fall (Eq. 15a). This
of proton generation and consumption, with implications for gives a value (Fig. 1A) much larger (Eq. 15b) than both the
the quantification of muscle buffer capacity. Their argument traditional physicochemical calculation (Eq. 9) and indeed
(21) is supported (in their Fig. 15) by the observation that larger than the traditional concept (the metabolic version, in the
efflux of protons during nonischemic exercise exceeds that of present terminology) of apparent buffer capacity (Eq. 15a;
lactate (15) and by calculations comparing the two approaches Fig. 1B).

Fig. 1. Application to ischemic exercise. Illustration of the application of the analyses in Table 2 to data from a 31P MRS study of ischemic forearm exercise
(16). It shows, for each analysis, the summed processes of proton generation and proton consumption: Here, these are equal by definition, as comparing in vivo
and ex vivo estimates of buffer capacity is not the point [see elsewhere for this, and for the technical issues (1, 16, 17)]. Using data from whole 3 min of exercise,
where ⌬pH ⬇ 0.5 and ⬃60% of ATP comes from glycogenolysis, shows the analysis of Robergs et al. (A) and the traditional analysis in both its versions (B);
the identical x-axes show how much larger proton generation/consumption terms is in the Robergs’ analysis. Various buffer capacities are shown next to the
analysis from which they derive, in units of mmol (l cell water)⫺1 (pH unit)⫺1, or slykes (23). (C) shows the results of both analyses for the first exercise data
point, where lactate generation has not begun and ⌬pH ⬇ ⫺0.02 (i.e., a small pH increase); note the different x-axis to (A) and (B), as the quantities are much
smaller. At this point the contribution of lactate to proton production (on the traditional view) and of pyruvate synthesis to proton production and pyruvate
reduction to proton consumption (21) are zero. On the traditional view the buffered and consumed components are equal and opposite (as pH rises), and the
glycolytic proton load is zero; the net proton load to the cytosolic buffers is therefore negative. The estimated buffer capacity (Eq. 9) is the negative ratio of the
(positive) consumed protons to the (negative) pH fall. In the Robergs analysis, total proton production is always positive, whether or not lactate accumulates (Eq.
13) in the absence of lactate accumulation, this proton load is solely from hydrolysis of the ATP generated via creatine kinase (Eq.s. 4 and 13), and such a positive
proton generation cannot explain the rise in pH; the total buffer capacity (Eq. 15a) therefore appears negative. In fact, this proton generation is exceeded by proton
consumption associated with flux through the creatine kinase reaction in the direction of phosphocreatine breakdown (Eq. 6b), the difference between the two
being the net proton consumption by the Lohmann reaction (Eq. 7), which in the traditional view readily explains the pH rise. The difference between the
traditional estimate of physicochemical ␤ at the end (B) and start of exercise (C) is due partly to the contribution of phosphate at the end of exercise and partly
to the effect of lower pH on both phosphate and the other passive buffers of the cell (16). Note that if the calculations in (A) used the values assumed by Robergs
et al. (Table 1A) rather than the correct stoichiometric coefficients derived from (17), the total quantity of protons and thus the total buffer capacity would be
little affected (only 6% higher); however, the implicit proportion of total proton production by ATP hydrolysis (Eq. 5a) would be 1.5 [as glycogenolysis to lactate
is a net consumer of protons at resting pH (21)] rather than the correct value of 1.1 decreasing to 0.6 as pH falls (Table 1A), and the overall fraction of total
proton production (Eq. 13) due to ATP hydrolysis would consequently appear to be 80%, twice the correct figure of 40%. The use of the approximate coefficients
misrepresents the balance between these components. An implication of the simplified Robergs stoichiometry assumptions is that the Lohmann reaction (Eq. 7)
consumes no protons, which would clearly have a large effect on (C, top).

AJP-Regul Integr Comp Physiol • VOL 289 • SEPTEMBER 2005 • www.ajpregu.org


Downloaded from journals.physiology.org/journal/ajpregu (210.050.069.126) on June 26, 2022.
Letters to the Editor
R900
However, by setting the total proton generation of Robergs phosphate concentration, and creatine kinase (16), insofar as
et al. (Eq. 13) equal to their total proton consumption (Eq. 14), this influences the relative proportion of protons buffered and
we obtain precisely the traditional expression (Eq. 8). Thus consumed. It has the technical advantage of being equal to the
whatever detailed calculations underlie Figs. 16 and 17 in true physicochemical buffer capacity in the hypothetical ab-
Robergs et al. (21), these calculations cannot influence whether sence of phosphocreatine breakdown but the disadvantage of
or not the implied value of cytosolic buffer capacity matches being zero in the absence of lactate production. This concept
that obtained by titration of muscle homogenate in vitro. might be seen as carrying useful information about metabolic
regulation (16), or else as fatally flawed (21), its experimental
COMPARISON OF THESE TWO APPROACHES
simplicity outweighed by conceptual obscurity.
The conceptual difference between these approaches is a mat- The relative proportion of protons buffered and consumed
ter of labeling the components of proton balance (compare features also in the relationship between total and physico-
Figs. 1A and 1B) but with real consequences for the concept of chemical buffer capacities (Eq. 15b) but in a way that makes
cellular buffer capacity. Robergs et al. take a broad view of the the total buffer capacity exactly twice the true buffer capacity
proton load, including all the protons produced by ATP hydro- in the absence of phosphocreatine breakdown and negative in
lysis (Eq. 4a) and by glycogenolysis to pyruvate (Eq. 1b), and the absence of lactate production. I suggest that total buffer
also of proton consumption, including the protons accounted capacity is no more useful than apparent buffer capacity and
for by reduction of pyruvate to lactate (Eq. 2) and by ATP physiologically more obscure.
synthesis at the expense of phosphocreatine (Eq. 6b). However,
CONCLUSION
much of this proton generation is cancelled by proton con-
sumption, so only a fraction remains to be buffered in the Robergs et al. have usefully called attention to some ambigu-
cytosol. ities and confusions in the literature on extra- and intracellular
The key point is what proton buffering means. To tradition- acidification by glycolysis. In the application to exercising
alists, it is physicochemical buffering (Eq. 9), sometimes muscle, however, they have reached a potentially misleading
expanded to include proton consumption in the Lohmann conclusion about muscle cell buffer capacity. The traditional
reaction (Eq. 7; 23) [although I prefer to distinguish these view criticized by Robergs et al. (21) certainly appears to
(16)]. The traditional view takes account of both the protons neglect the roles of lactate synthesis in consuming the protons
consumed by phosphorylation of ADP by phosphocreatine (Eq. formed earlier in glycogenolysis and of ATP hydrolysis as the
6) and the protons produced by the hydrolysis of the ATP main source of the protons that accompany lactate. However,
supplied at the expense of phosphocreatine (Eq. 4c) and com- their presentation, which ignores for simplicity the pH depen-
bines them into the net proton consumption by the Lohmann dence of the stoichiometry (6, 12, 14, 26), does not show how
reaction (Eq. 7). What remains after this is subtracted from the the importance of this point decreases as cell pH falls during
glycogenolytic proton load in the physicochemical version of exercise and that it does not in any case affect the logic of the
this approach is the net proton load, which is buffered in the traditional calculations. Additionally, Robergs et al. (21) make
cytosol (16); in the metabolic version, the ratio of total proton what is essentially a labeling error in the analysis of cellular
load to pH fall is a measure of the combined effect of physi- proton buffering. Focusing attention on the individual reac-
cochemical and metabolic buffering (16, 23), which may or tions, they do not take account of how some of the processes of
may not be useful (see below). In both versions of the tradi- proton generation are canceled out by those of proton con-
tional approach, metabolic buffering means the Lohmann re- sumption. Thus their total muscle buffer capacity does not
action. seem to identify a useful property of muscle physiology.
In the proposed new view, by contrast, the concept of
APPENDIX
metabolic buffering is expanded (21) to include two processes
of proton consumption: reduction of pyruvate to lactate (Eq. 2) There are some possible inconsistencies in Figs. 16 and 17 of
and ATP regeneration at the expense of phosphocreatine (Eq. Robergs et al. (21), and the relationship to the experimental
6b). However, these are partially cancelled out by two pro- papers cited (3, 19, 24, 25) is not entirely clear. For example,
cesses of proton generation: hydrolysis of ATP arising from the 125 mmol proton/kg muscle used as the denominator to
phosphocreatine breakdown (Eq. 4c) and glycolysis to pyru- estimate total muscle buffer capacity (Eq. 15 here) as 208
vate (Eq. 1b). Thus the ratio of this expanded total proton load slykes (evidently mmol/kg wet wt) appears inconsistent with
(Eq. 13) to pH fall, proposed as total muscle buffer capacity the ⬃100 mmol proton/kg muscle shown in Fig. 17. The 72
(21), has no obvious physiological interpretation (Eq. 15b). It mmol proton/kg muscle used to calculate the metabolic buff-
has a complicated relationship to physicochemical cytosolic ering of 120 slykes appears to be the sum of the components
buffer capacity: when acidification is substantial, total muscle labeled Pi, CrP and lactate in Fig. 17, which I take to be the
buffer capacity considerably exceeds physiochemical cytosolic protons buffered by the inorganic phosphate component of ␤
buffer capacity (Fig. 1A); in the special case of no lactate and the protons consumed by phosphocreatine hydrolysis
accumulation (Fig. 1C), which poses no problem to the tradi- (Eq. 6b) and the reduction of pyruvate to lactate (Eq. 2),
tional approach (4), total cellular buffering capacity is nega- respectively. However, the first of these should presumably
tive, a reductio ad absurdum. be considered structural or static buffering: even though
The apparent buffer capacity (Eq. 12a), which Robergs et al. phosphate increases during exercise, the proton-binding
criticize has similar defects: it too overestimates physicochem- effect of phosphate requires a fall in pH (to drive the ionic
ical buffer capacity, although in a more algebraically straight- equilibrium Pi2⫺ ⫹ H⫹ 7 HPi⫺ to the right) but no enzyme
forward way (Eq. 12b), which is related to the metabolically action; in this it differs from the indubitably metabolic
important relationship between glycogen phosphorylase flux, Lohmann reaction, which requires enzyme activity but not a
AJP-Regul Integr Comp Physiol • VOL 289 • SEPTEMBER 2005 • www.ajpregu.org
Downloaded from journals.physiology.org/journal/ajpregu (210.050.069.126) on June 26, 2022.
Letters to the Editor
R901
change in pH; indeed, it can consume protons when pH is 12. Hochachka PW and Mommsen TP. Protons and anaerobiosis. Science
rising (Fig. 1C) and the static buffers are giving protons up. 219: 1391–1397, 1983.
13. Hultman E and Sahlin K. Acid-base balance during exercise. Exerc
The remaining structural (physicochemical) buffer capacity Sport Sci Rev 8: 41–128, 1980.
of 88 (⫽208 ⫺ 120) slykes is not, as claimed, close to the 14. Jones NL. Hydrogen ion balance during exercise. Clin Sci (Lond) 59:
structural buffer capacity reported in (23); the 42 slykes they 85–91, 1980.
cite in Fig. 16 (21) as a value obtained by homogenate 15. Juel C, Klarskov C, Nielsen JJ, Krustrup P, Mohr M, and Bangsbo J.
Effect of high-intensity intermittent training on lactate and H⫹ release
titration appears to be a value predicted from measured and from human skeletal muscle. Am J Physiol Endocrinol Metab 286:
estimated cytosolic concentrations (23) and is anyway at the E245–E251, 2004.
low end of the published range of homogenate-titration 16. Kemp GJ, Roussel M, Bendahan D, Le Fur Y, and Cozzone PJ. Interre-
values (16, 23), which are often overestimates (1). Because lations of ATP synthesis and proton handling in ischaemic exercise studied by
31
P magnetic resonance spectroscopy. J Physiol 535: 901–928, 2001.
of these uncertainties, in this letter, I have taken a theoret- 17. Kushmerick MJ. Multiple equilibria of cations with metabolites in muscle
ical approach supported by illustrative data from another bioenergetics. Am J Physiol Cell Physiol 272: C1739 –C1747, 1997.
source. 18. Lambeth MJ and Kushmerick MJ. A computational model for glyco-
genolysis in skeletal muscle. Ann Biomed Eng 30: 808 – 827, 2002.
19. Medbo JI and Tabata I. Anaerobic energy release in working muscle
REFERENCES during 30 s to 3 min of exhausting bicycling. J Appl Physiol 75:
1. Adams GR, Foley JM, and Meyer RA. Muscle buffer capacity estimated 1654 –1660, 1993.
from pH changes during rest-to-work transitions. J Appl Physiol 69: 20. Newcomer BR and Boska MD. Adenosine triphosphate production rates,
968 –972, 1990. metabolic economy calculations, pH, phosphomonoesters, phosphodi-
2. Arthur PG, West TG, and Hochachka PW. Quantitative modelling of esters, and force output during short-duration maximal isometric plantar
the effects of selected intracellular metabolites on pH in fish white muscle. flexion exercises and repeated maximal isometric plantar flexion exercises.
J Exp Biol 200: 1189 –1200, 1997. Muscle Nerve 20: 336 –346, 1997.
3. Bangsbo J, Gollnick PD, Graham TE, Juel C, Kiens B, Mizuno M, and 21. Robergs RA, Ghiasvand F, and Parker D. Biochemistry of exercise-
Saltin B. Anaerobic energy production and O2 deficit-debt relationship induced metabolic acidosis. Am J Physiol Regul Integr Comp Physiol 287:
R502–R516, 2004.
during exhaustive exercise in humans. J Physiol 422: 539 –559, 1990.
22. Russ DW, Elliott MA, Vandenborne K, Walter GA, and Binder-
4. Bendahan D, Kemp G, Roussel M, Le Fur Y, and Cozzone P. ATP
Macleod S. A Metabolic costs of isometric force generation and mainte-
synthesis and proton handling in short periods of exercise and subsequent
nance of human skeletal muscle. Am J Physiol Endocrinol Metab 282:
recovery. J Appl Physiol 94: 2391–2397, 2003.
E448 –E457, 2002.
5. Blei ML, Conley KE, and Kushmerick MJ. Separate measures of ATP
23. Sahlin K. Intracellular pH and energy metabolism in skeletal muscle of
utilization and recovery in human skeletal muscle [correction in J Physiol
man. Acta Physiol Scand Suppl 455: 1–56, 1978.
475: 548, 1994]. J Physiol 465: 203–222, 1993.
24. Spriet LL, Soderlund K, Bergstrom M, and Hultman E. Anaerobic
6. Busa WB and Nuccitelli R. Metabolic regulation via intracellular pH.
energy release in skeletal muscle during electrical stimulation in men.
Am J Physiol Regul Integr Comp Physiol 246: R409 –R438, 1984.
J Appl Physiol 62: 611– 615, 1987.
7. Conley KE, Blei ML, Richards TL, Kushmerick MJ, and Jubrias SA.
25. Spriet LL, Soderlund K, Bergstrom M, and Hultman E. Skeletal
Activation of glycolysis in human muscle in vivo [correction in Am J muscle glycogenolysis, glycolysis, and pH during electrical stimulation in
Physiol Cell Physiol 276: Ca1, 1999]. Am J Physiol Cell Physiol 273: men. J Physiol 62: 616 – 621, 1987.
C306 –C315, 1997. 26. Wolfe CL, Gilbert HF, Brindle KM, and Radda GK. Determination of
8. Conley KE, Kushmerick MJ, and Jubrias SA. Glycolysis is indepen- buffering capacity of rat myocardium during ischemia. Biochim Biophys
dent of oxygenation state in stimulated human skeletal muscle in vivo. Acta 971: 9 –20, 1988.
J Physiol 511: 935–945, 1998.
9. Connett R. Analysis of metabolic control: new insights using a scaled
creatine model. Am J Physiol Regul Integr Comp Physiol 254: R949 – Graham Kemp
R959, 1988. Division of Metabolic and Cellular Medicine,
10. Crowther GJ, Milstein JM, Jubrias SA, Kushmerick MJ, Gronka RK, School of Clinical Sciences
and Conley KE. Altered energetic properties in skeletal muscle of men Faculty of Medicine,
with well-controlled insulin-dependent (type 1) diabetes. Am J Physiol
Endocrinol Metab 284: E655–E662, 2003.
University of Liverpool,
11. Dennis SC, Gevers W, and Opie LH. Protons in ischemia: where do they Liverpool L69 3GA, UK
come from; where do they go to? J Mol Cell Cardiol 23: 1077–1086, 1991. gkemp@liv.ac.uk

AJP-Regul Integr Comp Physiol • VOL 289 • SEPTEMBER 2005 • www.ajpregu.org


Downloaded from journals.physiology.org/journal/ajpregu (210.050.069.126) on June 26, 2022.

You might also like