Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Biophys J. 2023 Feb 7; 122(3): 544–553.
Published online 2022 Dec 22. doi: 10.1016/j.bpj.2022.12.028
PMCID: PMC9941725
PMID: 36564946

Bridging ribosomal synthesis to cell growth through the lens of kinetics

Associated Data

Supplementary Materials

Abstract

Understanding prokaryotic cell growth requires a multiscale modeling framework from the kinetics perspective. The detailed kinetics pathway of ribosomes exhibits features beyond the scope of the classical Hopfield kinetics model. The complexity of the molecular responses to various nutrient conditions poses additional challenge to elucidate the cell growth. Herein, a kinetics framework is developed to bridge ribosomal synthesis to cell growth. For the ribosomal synthesis kinetics, the competitive binding between cognate and near-cognate tRNAs for ribosomes can be modulated by Mg2+. This results in distinct patterns of the speed – accuracy relation comprising “trade-off” and “competition” regimes. Furthermore, the cell growth rate is optimized by varying the characteristics of ribosomal synthesis through cellular responses to different nutrient conditions. In this scenario, cellular responses to nutrient conditions manifest by two quadratic scaling relations: one for nutrient flux versus cell mass, the other for ribosomal number versus growth rate. Both are in quantitative agreement with experimental measurements.

Significance

Ribosomes play essential roles in protein synthesis which underpins cell growth. Mg2+-dependent ribosomal synthesis performance in terms of speed – accuracy relation is illustrated by the generalized Hopfield model. Integrating the characteristics of ribosomal synthesis into cell growth model yields adaptive cell growth kinetics through Mg2+-dependent responses to different nutrient conditions.

Introduction

Protein synthesis is one of the most resource-intensive processes that occurs during bacterial growth. Since ribosomes are responsible for polypeptide synthesis, studying ribosomal kinetics across multiscale is important because the molecular performances will ultimately affect the cellular behaviors. At the molecular level, protein synthesis has the elongation speed of 10∼22 amino acids per second (1,2,3) and the accuracy of one mistake per 3×103 codons (4,5,6), enabling the rapid growth to prevail over rivals for high fitness (7). Studying the ribosomal kinetics in terms of speed – accuracy relation is thus necessary to understand the performance of polypeptide synthesis (8,9,10). At the cellular level, cells adopt several different strategies for optimal growth in response to nutrient conditions (11). Cells can increase protein synthesis speed for rapid growth, manifesting in the Michaelis-Menten relation between protein synthesis speed and growth rate (12). The variation of protein synthesis accuracy was observed to adjust cell growth (13,14), suggesting the responses on cellular activities from the ribosomal speed and accuracy relation. (15,16).

The kinetics pathway of ribosomes comprises two stages: initial selection and proofreading (17). At initial selection, ribosomes select cognates over near-cognates with significantly faster forward rate constants owing to the correct codon-anticodon interactions. In case near-cognates pass the initial selection, ribosomes can discard near-cognates at proofreading energized from the irreversible guanosine-5′-triphosphate (GTP) hydrolysis (18,19,20). The cognates are clearly favored during the peptide bond formation within the ribosomes which invoke the delicate stepwise “proton shuttle” process. (21). The in vitro discrimination factor of selecting cognates over near-cognates is 30∼60 for initial selection (4,19,22) and 20∼100 for proofreading (4,19,20,23). The product of the discrimination factor of the two stages agree with the in vivo accuracy of 3×103 (4,5).

Ribosomes exhibit the trade-off relation between speed and accuracy during peptide synthesis. The mutation in ribosomes contains an asparagine instead of lysine at position 42 of protein S12, enabling ribosomes to achieve better accuracy in protein production by reducing the overall elongation speed for cognates and near-cognates (24). Alternatively, increasing Mg2+ enables ribosomes to produce proteins more rapidly but with less accuracy because the speed of near-cognates accelerates at a faster rate with respect to Mg2+ than that of the cognates (15,16). The acceleration of near-cognates is attributed to less dissociation of the complexes between ribosomes and near-cognates, thus stabilizing near-cognates’ binding to ribosomes and causing more competition for cognates selection (25). Hopfield model can explain the trade-off relation by suggesting that synthesis speed increases and accuracy reduces when rejection rate constants decrease (18). Recently, ribosomes have been shown to discriminate near-cognates against cognates with sluggish forward speed (e.g., the net speed of initial selection is 1.6 and 100 μM1s1 for near-cognates and cognates, respectively) (19,20). Moreover, less rejection can lead to enhanced competition with more ribosomes occupying near-cognates, especially at high Mg2+ due to diminishing contribution of proofreading to total accuracy (16,26). Consequently, near-cognates can slow down the translation speed once they bind to ribosomes. Insides cells, ribosomes frequently encounter near-cognates because of their much higher number compared with cognates (20,27). However, it has remained unclear how the kinetics difference between cognates and near-cognates influences ribosomal partition between the corresponding two complexes and the peptide synthesis speed – accuracy relation.

Cellular adaptation to nutrient fluxes manifests in a linear relation between ribosomal mass fraction and growth rate when bacteria reach exponential, steady-state growth at different nutrient quality (28,29,30,31). The linear relation emerges from the optimal proteomic arrangement for the fastest growth (32). The effects of protein synthesis accuracy on cell growth were extensively investigated. The growth rate of E. coli cells decreased by 26% due to a twofold decrease in the accuracy caused by wrong isoleucine incorporation into the poly-phenylalanine chain (33). The decrease in accuracy is attributed to mutations at ribosomal proteins S4 and S5 disabling termination of protein synthesis at stop codons (i.e., readthrough of stop codon) (33). Readthrough of stop codon is also induced by mutations tufA8 and tufB103 in elongation factor Tu (34). The mutations caused a fourfold increase in misincorporation of leucine into the poly-phenylalanine chain, consequently reducing the growth rate of S. typhimurium cells by nearly 30% (34). Besides, cells have signaling pathways to maintain intracellular Mg2+ in response to perturbations amid the cell growth (35,36,37). If Mg2+ import is enhanced due to overexpression of the MgtE gene (35,36), intracellular Mg2+ increases and causes both accuracy and cell growth to reduce. Increasing Mg2+ from 4.5 to 14 mM caused a nearly tenfold increase of leucine misincorporation into the poly-phenylalanine chain (38), and further raising Mg2+ to 1.25 M reduced growth rate of E. coli cells sixfold (39).

In this article, we address the speed – accuracy relation of ribosomes and its connection to cell growth. Hopfield model is extended into generalized Hopfield model (GHM) to account for the in vivo conditions where the competition between cognates and near-cognates requires ribosomes to distinguish the two species with different kinetics. The GHM shows that Mg2+ modulates the ribosomal speed – accuracy relation from trade-off to positive correlation. We then study the effects of ribosomal speed – accuracy relation on cell growth via a kinetics coupled bacterial growth model (BGM). Cellular response toward nutrient conditions can be regulated by Mg2+ which enables cells to optimize growth rate.

Materials and methods

The kinetics scheme of ribosomes for peptide bond formation is elaborated. The free energy landscape analysis is invoked for the study of Mg2+-dependent rate constants, thereby capturing the overall trend of ribosomal speed – accuracy as a function of Mg2+. The results of GHM are integrated into the BGM, as illustrated in Fig. S1, to describe how cells control Mg2+ to modulate accuracy and speed for the optimal growth in response to nutrient conditions. The codes are available at https://github.com/haibinsu/Ribosome_cell_growth.

Protein synthesis from the GHM

The GHM includes three key states of ribosome: free ribosomes (Rfree), GTP-activated state of initial selection (RAct), and hydrolyzed ternary complexes involved proofreading (RPR), as shown in Fig. 1 a. Initially, Rfree waits for the binding of ternary complexes T3 to form RAct with bimolecular rate constants kfc/nc. The superscripts c,nc denote cognates and near-cognates, respectively. The kf corresponds to the overall measured speed (denoted as kcat/KM) of initial selection consisting of binding between ternary complexes and ribosomes, codon recognition, and GTPase activation (Fig. S2 a; Eq. S19). Once ribosomes reach RAct, they quickly enter RPR due to rapid hydrolysis of GTP with rate constant khyd. At proofreading, ribosomes can either form a peptide bond or reject hydrolyzed ternary complexes with rate constants kpepfc/nc and krejc/nc, respectively. If peptide formation occurs, ribosomes translocate to the next codon of messenger RNA, become Rfree, and start the next round of peptide synthesis. If rejection happens, which is most likely for near-cognates, ribosomes become Rfree and repeat the selection process until a peptide bond is formed. The two scenarios indicate that ribosomes eventually return to Rfree to wait for T3. The three states form a cycle where ribosomes continuously synthesize protein by selecting cognates, which are outnumbered by the pool of near-cognates. The selection of cognates under the competition of near-cognates is represented as the two reaction pathways with different kinetics as shown in Fig. 1 a.

An external file that holds a picture, illustration, etc.
Object name is gr1.jpg

The generalized Hopfield model (GHM) of ribosomal protein synthesis. (a) Three states are: free ribosomes (Rfree), GTPase activation at initial selection (RAct), and proofreading (RPR) comprising peptide formation and rejection. Arrow thickness reflects the magnitude of rate constants depending on cognates or near-cognates. (b) The partition of ribosomes in three states at various Mg2+ is dominated by Rfree (white), RPRc for cognates (green), and RPRnc for near-cognates (brown). (c) The relation among ksyneff, accuracy, and Mg2+. The magenta and blue dots indicate the maximum value of ksyneff. To see this figure in color, go online.

Based on Fig. 1 a, the kinetics of ribosomes is as follows:

d[Rfree]dt=(krejc+kpepfc)[RPRc]+(krejnc+kpepfnc)[RPRnc](kfcfc+kfncfnc)[T3][Rfree]+kbc[RActc]+kbnc[RActnc],
(1a)
d[RActc]dt=kfcfc[T3][Rfree](kbc+khyd)[RActc],
(1b)
d[RActnc]dt=kfncfnc[T3][Rfree](kbnc+khyd)[RActnc],
(1c)
d[RPRc]dt=khyd[RActc](krejc+kpepfc)[RPRc],
(1d)

and

d[RPRnc]dt=khyd[RActnc](krejnc+kpepfnc)[RPRnc].
(1e)

The total concentration of ribosomes [Rt] is 10 μM, and it is assumed constant because of enzyme conservation (i.e., the sum of (1a), (1b), (1c), (1d), (1e) is zero). [Rt] is partitioned into five chemical states during amino acid incorporation (Fig. 1 a):

[Rfree]+[RActc]+[RActnc]+[RPRc]+[RPRnc]=[Rt].
(1f)

[T3]=100μM, while the number fraction of cognate and near-cognate species are fc=0.02 and fnc=0.15, respectively. All values correspond to in vivo conditions (20,27,40). The values of rate constants kfc and kfnc are from the in vitro data measured at different Mg2+ (16,41). The backward rate constants kbc/nc are neglected because they are much smaller than the forward rate constants kfc/nc. The remaining rate constants are given in Table S1. Eqs. 1a–1e are solved at steady state where there are no changes in concentrations with time. Eq. 1f acts as a constraint since there are only four independent equations among Eqs. 1a–1e.

The influence of Mg2+ on the protein synthesis speed – accuracy relation is studied by using the GHM in conjunction with the experimental data (16,41). In particular, the bimolecular rate constant kf corresponds to the kcat/KM of initial selection (Eq. S19), and the net rate constant from Rfree to RPR is the kcat/KM of overall peptide formation (Table S1; supporting material section 2.2). Expressions of kfc and kfnc are based on the free energy landscape of ribosomes at initial selection (supporting material section 2.3). The expressions show good agreement with the experimental data as depicted in Fig. S2 b, thereby capturing the rate constants’ dependence on Mg2+.

Cell growth from the BGM

Based on Fig. 2 a, the kinetics of the BGM is expressed as follows:

M=mRR+mPP+a˜=ρV,
(2a)
dRdt=uRmRR·m(A)ksynmaxcell(A)a˜KMa˜(A)NAV+a˜kdR,
(2b)
dPdt=1uRmPR·m(A)ksynmaxcell(A)a˜KMa˜(A)NAV+a˜kdP,
(2c)

and

da˜dt=Pkmeteff1+(a˜KeNAV)2R·m(A)ksynmaxcell(A)a˜KMa˜(A)NAV+a˜Jmtn,
(2d)
where

R, P, and a˜ denote the number of ribosomes, metabolic proteins, and amino acids, respectively. mR and mP are the masses of ribosome and metabolic protein, respectively, with all masses expressed in terms of the number of amino acids (Table S2). The three components approximately contribute to total cell mass M (Table S2). The approximation follows the experimental results showing that protein and RNA dry mass can take up to 70% of cell dry mass (44). Protein mass density is approximately constant across different growth rates (45,46) as shown in Fig. S6 b. The Avogadro constant NA is used to express concentrations in term of molecular numbers. Protein synthesis flux depends on accuracy A via accuracy dependent KMa˜(A), ksynmaxcell(A), and m(A). The KMa˜ of amino acid is presumably proportional to the KMeff of the ternary complex (Eqs. S47 and S48; Table S2). The maximal protein synthesis speed ksynmaxcell also couples with accuracy in the framework of GHM. The modulation factor m can capture the effect of accuracy in the cells. The dependence of m on accuracy across various nutrient conditions is illustrated in Fig. S7 a. Ribosomes synthesize themselves with allocation uR, suggesting that the self-catalytic activity of ribosomes is the origin of exponential growth (29,32). The remaining allocation of protein synthesis flux is for metabolic proteins that supply amino acids (29,32). Cells limit the number of amino acids to avoid overproduction by negative feedback via inhibition constant Ke (32,47). For moderate-fast growth, the maintenance flux Jmtn can be neglected because most of the nutrient intake flux is allocated for growth (48). We choose the values of all variables and expressions so that the output parameters, such as growth rates and the number and mass fraction of ribosomes, are within the range of experimental data. The physical meanings and values of remaining parameters are further provided in Tables S2, S5, and S6.

An external file that holds a picture, illustration, etc.
Object name is gr2.jpg

Cellular characteristics by the bacterial growth model (BGM). (a) The schematic illustration of BGM (b) The characteristics of ribosomal synthesis in terms of protein synthesis speed ksyneffcell-accuracy relation (at [T3]=100μM). Experimental data are from refs. 16 and 41. (c) Mg2+ regulated optimal cell growth rate under different nutrient conditions. The optimal growth rates λ as large, solid dots. (d) The quadratic scaling between number of ribosomes per cell R and optimal growth rate λ with the experimental data of E. coli (1,42). The inset shows a linear relation between λ and ribosomal mass fraction φR with the experimental data of E. coli (3,29,31). (e) The quadratic scaling between intake flux JM and cell mass M with the experimental data of E. coli (1,30,42). The right inset shows metabolic fluxes of different bacteria species (43). The linear Mλ relation is shown in the left inset. To see this figure in color, go online.

The important parameters in the BGM are amino acid-production rate constant kmeteff, cell mass M, and accuracy A, as well as accuracy dependent maximal speed ksynmaxcell, Michaelis constant KMa˜, and modulation factor m. The values of kmeteff, listed in Table S5, fall within the experimental range for various substrates summarized in Table S3. The dependence of kmeteff on nutrient conditions is captured through its increasing values indicating better nutrient quality (32). The values of cell mass M, shown in Table S5, are physically consistent with the experimental values in Table S4. Besides, M presumably follows growth rate in a linear fashion. The presumption is based on the experimental results shown in Fig. S6 a (30) and the left inset of Fig. 2 e (1). For bacterial growth, the maximal value of ksynmaxcell is 22 s1 in order to match in vivo condition by changing kpepfc to 22 s1. The values of ribosomal rate constants, shown in Fig. S9 and Table S6, follow the experimental values summarized in Table S1. The values of accuracy, ksynmaxcell, and KMa˜ are calculated based on Tables S2 and S6 and Eqs. 3b, 4b, and 4c, respectively.

Results and discussion

The GHM extends Hopfield model by considering the kinetics difference between ternary complexes and their competition for ribosomes, thereby establishing that the coupling between protein synthesis speed and accuracy is the basis for the speed – accuracy relation. The GHM also establishes quantitatively the modulation of Mg2+ on ribosomal speed – accuracy relation. Overall, the speed – accuracy relation subject to Mg2+ regulation is incorporated in the BGM to demonstrate that nutrient quality dictates the cellular regulation of Mg2+ intake flux to optimize both accuracy and ribosomal content for optimal growth.

The protein synthesis

The GHM is an extension of Hopfield model to address Mg2+ effect on protein synthesis speed – accuracy relation. Like Hopfield model (18,49,50), the protein synthesis network in the GHM has three chemical states: an available ribosomal state waiting for ternary complexes and two other states where ribosomes are occupied with tRNA (49,50) (e.g., RAct and RPR in the GHM). The transition between two tRNA-occupied states is driven by the irreversible hydrolysis enabling ribosomes to reject near-cognates at proofreading. The rate constants kf,khyd, and kpepf are independent of ternary complexes in Hopfield model as depicted in Fig. S3 (18). In the GHM, these rate constants are based on biochemical measurements, for example, kfc100μM1s1 and kfnc1.6μM1s1 (19,20). Peptide formation rate constant of cognates is also faster than that of near-cognates (i.e., kpepfc is 7∼22 s1 (10), whereas kpepfnc is 0.3 s1 (19)). Moreover, near-cognates face higher rejection due to much larger krejnc than krejc of cognates, which is less than 0.1 s1 (19). The kinetics difference between cognates and near-cognates is also considered in previous work (49,50). The kf in those models corresponds to the codon recognition step (24); the values of kf are first-order rate constants. In contrast, the kf in the GHM is the bimolecular rate constant, which corresponds to the net rate of ribosomes transitioning from the free to the GTPase-activated state (Eq. S19). The competition between cognates and near-cognates for ribosomes depends on not only the bimolecular rate constants kfc and kfnc but also the concentrations of the two ternary complexes. The GHM considers the concentration difference via the number fraction of cognates fc and near-cognates fnc with respect to total [T3]. The values of fc and fnc are in good agreement with in vivo measurements (20,27,40).

The steady-state peptide production fluxes Jt and accuracy A are defined as follows:

Jt=Jc+Jnc=[RPRc]kpepfc+[RPRnc]kpepfnc
(3a)

and

A=JcJnc=[RPRc]kpepfc[RPRnc]kpepfnc.
(3b)

The accuracy A, defined as the peptide flux ratio between cognates and near-cognates in Eq. 3b, comprises the contribution from initial selection and proofreading (Eq. S10), as shown by in vivo measurements (18,40). The definition also takes into account the concentration difference between cognates and near-cognates. The protein synthesis speed ksyneff can be further expressed in the form of Michaelis-Menten kinetics (supporting material section 2.2):

ksyneff=Jt[Rt]=ksynmax[T3]KMeff+[T3],
(4a)

where ksynmax and KMeff are approximated as follows (Eqs. S8 and S9):

ksynmaxfckfc(fckfckpepfc+fnckfnckrejnc)1
(4b)

and

KMeff(fckfckpepfc+fnckfnckrejnc)1.
(4c)

ksynmax and KMeff represent the maximum protein synthesis speed and Michaelis constant (i.e., indicating ternary complex concentration [T3] that gives half of ksynmax), respectively. Both ksynmax and KMeff depend on the competition between cognates and near-cognates through their number fraction fc and fnc. Besides, ksynmax and KMeff can vary because some rate constants (e.g., kfc and kfnc) depend on Mg2+. More importantly, changes in the rate constants of some ribosomal states affect ksynmax, KMeff, and accuracy A simultaneously (Eqs. 3b and S10), suggesting the intimate interplay among the three quantities. The dependence of the speed – accuracy relation on Mg2+ is elaborated in following subsection.

[Mg2+] effect on protein synthesis

The partition of ribosomes at different Mg2+ relates to their Mg2+-dependent rate constants. The ribosomal partition provides molecular insight on how the allocation of ribosomes in the network effectively manifests into ksyneff and accuracy. The three dominant states are Rfree, RPRnc, and RPRc, which are approximated as follows (for Mg2+ within 1∼8 mM):

[Rfree][Rt]r0exp(a0[Mg2+]),
(5a)
[RPRnc][Rt]rprncexp(aprnc[Mg2+]),
(5b)

and

[RPRc][Rt][1r0exp(a0[Mg2+])rprncexp(aprnc[Mg2+])],
(5c)

where r0=0.082,a0=0.3mM1,rprnc=0.002,andaprnc=0.62mM1 (Eqs. S40a–S40c). The approximation captures the trend of the data calculated based on the experimental rate constants and the GHM (Figs. S2 d and S4 a). Fig. 1 b shows the ribosomal percentage distribution of the three states at different Mg2+. The total percentage of RPRc and RPRnc rose and reached plateau, while Rfree plummeted as Mg2+ increased. At low Mg2+, RPRc dominated and occupied almost 95% at 3 mM because ribosomes prefer cognates to near-cognates. Beyond 3 mM, RPRnc accumulated, while the concentrations of RPRc and Rfree reduced. The dependence of ribosomal partition on Mg2+ is closely related to the modulation of Mg2+ on ksyneff and accuracy. The relation between ksyneff, accuracy, and Mg2+ was depicted as the three-dimensional curve (3D curve) shown in Fig. 1 c. When the 3D curve was projected onto the ksyneff-Mg2+ plane, ksyneff fell into “trade-off” and “competition” regimes at low (1∼3 mM) and high Mg2+ (3∼8 mM), respectively.

The “trade-off” regime exhibits the trade-off with increasing ksyneff and reducing accuracy. When Mg2+ increased from 1 to 3 mM, ksyneff increased and reached its maximum value. The increase of ksyneff is attributed to KMeff, which showed a nearly threefold decrease (Fig. S4 b), while ksynmax remained almost constant within 1∼3 mM. The increase also relates to the available 6% of Rfree (Fig. 1 b) serving as an extra surplus to associate with cognates and near-cognates. Rfree can bind more frequently to both cognates and near-cognates because kfc and kfnc increase, but the majority of Rfree associates with cognates and converts into RPRc due to much faster kfc. However, the accuracy experiences an almost sixfold reduction due to more ribosomes associating with near-cognates. The enhanced association of near-cognates closely relates to kfnc that increases more than kfc with increasing Mg2+ (Fig. S2 b; Table S1), leading to the trade-off curve as shown in the ksyneff-accuracy plane.

The “competition” regime shows the positive correlation between ksyneff and accuracy. ksyneff reduced quickly because of less formation at RPRc and more accumulation at RPRnc. The reduction indicates the enhanced association of near-cognates for limited Rfree. Near-cognates with increasing kfnc can rapidly associate with free ribosomes to form RPRnc with less rejection due to decreasing krejnc (Table S1; Fig. S2, b and c). Consequently, the concentration of RPRnc increases (Fig. 1 b) and accuracy reduces (Fig. 1 c). At the proofreading stage, sluggish peptide formation and little rejection increase the number of ribosomes binding to hydrolyzed near-cognates. The near-cognates-bound ribosomes are prohibited from being released to bind to cognates, causing drastic reduction of ksynmax and a modest decrease of KMeff (Fig. S4 b). The overall consequence is the positive correlation between ksyneff and accuracy when Mg2+ increased from 3 to 8 mM as indicated in the ksyneff-accuracy plane. The “competition” regime was also studied by single ribosome stochastic modelling (49). The optimal ksyneff is located at the transition from “trade-off” to “competition” at [Mg2+] of nearly 3 mM (Figs. 1 c and S4 a), which is within the measured range of Mg2+ inside E. coli (1∼5 mM) (16,36,37). Besides, the maximal speed can be maintained for Mg2+ within 1∼3 mM (supporting material section 2.3). Coincidentally, the in vitro globin synthesis by ribosomes in mice also showed a similar transition with our result (51) (supporting material section 3.5).

Two aspects of ribosomal kinetics can be further incorporated in GHM. Firstly, the in vivo protein synthesis shows that most of the errors fall into the error hotspots of some types of codons (e.g., mismatch of codons GGA and GAU to tRNAUUCGlu with error frequencies of 35×104 (52)). The error frequencies are two orders of magnitude higher than other codons (52). The effects of codons on peptide bond formation kinetics and accuracy need to be taken into account in GHM for studying the variation of error frequency. Secondly, considering that high Mg2+ drastically slowed down the mRNA translocation rate (53), the translocation step can be incorporated into GHM by partitioning the total protein synthesis time into amino acid incorporation and translocation for better characterization of ribosomal protein synthesis kinetics.

The bacterial growth

Cell growth depends on cellular regulation of Mg2+ in response to the environment. Cellular regulation of Mg2+ has been observed under two conditions: regular (2 mM) and low (0.1 μM∼1 mM) concentration of magnesium ions in the media. Under regular Mg2+ condition, B. subtilis cells modulated Mg2+ intake flux to cope with the antibiotic spectinomycin (54) and osmotic stress (55). Exposure to a sublethal amount of spectinomycin (2 mg/L) reduced protein synthesis speed and cell growth (54). Cells increased their growth rate by increasing Mg2+ intake flux so that Mg2+ can stabilize ribosomal structure and restore protein synthesis speed (54). Under osmotic stress induced by 1.2 M NaCl, cells restored growth rate by increasing Mg2+ intake to compensate for the loss of Mg2+ ions (supporting material section 3.5) (55). Under low Mg2+, both S. typhimurium and E. coli cells slowed down protein synthesis and used regulatory pathways to import Mg2+ ions (supporting material section 3.5) (56,57). Cell growth can be enhanced by adding Mg2+. For example, S. typhimurium wild-type cells had larger colonies on agar when Mg2+ in the agar increased from 40 to 500 μM (58). E. coli cells exhibited faster growth rate when extracellular Mg2+ increased within the range of 0.1 μM to 1 mM (59). In this work, we integrate Mg2+-dependent ribosomal speed – accuracy relation into the cell growth process analyzed by BGM.

Accuracy effect on cell growth

Cellular modulation of Mg2+ is represented by the speed – accuracy relation and modulation factor m. The effective protein synthesis speed ksyneffcell couples with accuracy because ksyneffcell is expressed in the Michaelis-Menten kinetics form consisting of ksynmaxcell and KMeff, as depicted in Fig. 2b. The coupling is also controlled by Mg2+ with each Mg2+ corresponding to a particular set of values of ksyneffcell and accuracy. As Mg2+ increased, ksyneffcell rose and fell while accuracy kept reducing as shown in Fig. 2 b. The Mg2+-dependent protein synthesis speed – accuracy relation thereby suggests that cells can regulate Mg2+ to tune accuracy for the fastest speed and the maximum growth rate. The sigmoid function of accuracy with the maximum value of one is used in the accuracy function m (Fig. S7 a) that reflects cellular regulatory network with the accuracy threshold below which cells cannot tolerate. Ribosomes with low accuracy produce nonfunctional proteins (6,60) and slow down growth rate (34,61). The functionality of some proteins is compromised by wrong amino acid incorporation (40). The probability of producing N proteins with at least T error-free proteins is a cumulative binomial distribution that behaves similarly to sigmoid function (62). The BGM also encodes the nutrient dependence into m to reflect that better nutrient qualities lead to the enhancement of accuracy and protein quality (Fig. S7 a).

For a given nutrient quality, cells optimize growth rate by adjusting proteomic content and accuracy (14,32). In the BGM, this optimization is achieved by the following two steps (supporting material section 3.2). Firstly, the growth rate is maximized locally at a given accuracy by tuning ribosomal mass fraction (32) (Fig. S8). Secondly, the maximum values of locally optimized growth rates will be taken as the optimal growth rates. All the maximal growth rates are depicted in Fig. 2 c as open dots, and the optimal among them at a given nutrient condition as solid dots represented as λ. As nutrient quality improves (e.g., increasing glucose concentration), cells can regulate Mg2+ intake fluxes for the optimal growth rates to occur within the typical intracellular Mg2+ of 1∼5 mM as observed in bacteria (16,36,37). Within this range, growth rate exhibits trade-off with accuracy as depicted in Fig. 2 c. Cells grow slowly with high accuracy under poor nutrient quality, while a rich nutrient condition enables cells to sacrifice accuracy for fast growth (14). Food quality thereby can trigger cells to regulate the Mg2+ intake flux to maximize growth rate in a similar manner to the aforementioned behaviors under stresses (e.g., antibiotics, osmotic difference, Mg2+ shortage).

Improvement in nutrient quality enables cells to achieve fast growth rate at the expense of accuracy, as shown in Fig. 2 c. Our analysis is consistent with (14), which also suggests that cells need to ensure the majority of proteins are functional to provide amino acids by maintaining slow and accurate protein synthesis when growth is hampered. When there is food abundance, cells divert their efforts to protein synthesis for rapid growth. By producing more ribosomes, cells need to ensure that all ribosomes work near their maximal speed ksynmaxcell by reducing the effect of proofreading to lower accuracy for fast growth (14). Besides, the BGM developed here can explain the positive correlation between growth rate and accuracy. Accuracy is considered at the cellular level comprising not only the performance of ribosomes but also cellular regulatory networks to maintain functional proteins. Therefore, the accuracy function depends on cellular responses toward external conditions. These conditions can perturb the relation between growth rate and accuracy, causing growth rate to positively correlate with accuracy. For example, mutations in elongation factor Tu caused a nearly fourfold decrease of accuracy and almost one-third reduction of growth rate (34). Raising the antibiotic streptomycin concentration from 2 to 10 μM increased the misincorporation of two amino acids nearly tenfold and decreased cell density almost threefold (63). Increasing Mg2+ from 4.5 to 14 mM reduced accuracy almost tenfold and decreased the growth rate sixfold when further raising Mg2+ to 1.25 M (38,39).

Cellular properties in the high accuracy regime

Under the normal growth conditions, the average accuracy of protein synthesis is around 3×103 (4,5), where growth rate and accuracy exhibit a trade-off scenario, as shown in Fig. 2 c. The majority of proteins are functional with minimal cellular control as modulated by the accuracy function. The values of ksynmaxcell are approximately constant across different optimal accuracy values (Fig. 2, b and c), as shown by the linear relation between ribosomal mass fraction and growth rate (inset of Fig. 2 d; Eq. S69).

Two quadratic relations emerge from the optimal growth rates λ and the corresponding cellular properties. The first relation is the number of ribosomes per cell R proportional to the square of growth rates λ, i.e., R(λ)2. Fig. 2 d shows that the quadratic relation captures well the monotonous increasing trend of growth rate as a function of the number of ribosomes. The second relation is the nutrient intake flux following a quadratic relationship with cell mass, i.e., JM(M)2. Both the theoretical and experimental fluxes were in good agreement with each other (1,30,42), as depicted in Fig. 2 e. The experimental protein synthesis fluxes are the product of the number of ribosomes per cell and protein synthesis speed (Table S4; supporting material section 3.1). Moreover, the theoretical values of JMM relation had a slope of two in log-log scale, and the slope is similar to the slope of 1.96 in (43) showing the scaling between metabolic flux and cell mass across bacteria species in the right inset of Fig. 2 e. The nearly quadratic scaling was presumably related to the scaling between genome size and cell mass with the power of 0.35 (43). Our analysis suggests another reason can be the scaling between ribosomes and cell mass across bacteria species. The quadratic scaling JM(M)2 can also indicate the efficiency of utilizing nutrient intake flux for cell growth (supporting material section 3.4.2). The biomass conversion figure of merit Z, defined as λ/JM, reduced with growth rate in the region of moderate to fast growth (Fig. S7 b), while it increased with growth rate in slow growth region (Fig. S11 b).

The coefficient involved in these two quadratic scaling relations can be written as: R=cFλ2/ksynmaxcell and JM=(M)2/cF. (Eqs. S53 and S54, respectively). The first quadratic relation between ribosomes number and growth rate originates from the two linear relations: cell mass versus growth rate (Fig. S6 a and the left inset of Fig. 2 e) and ribosomal mass fraction versus growth rate (the inset of Fig. 2 d). Since nutrient intake flux equals the product of the protein synthesis speed and the number of ribosomes, this results in the second quadratic relation of nutrient intake flux versus cell mass. The quadratic scaling equations captured the experimental data well due to the common factor cF of 2.15×109 (aa·h), as shown in Fig. S12. The common factor cF indicates the importance of ribosomes and connects various cellular properties: the number of ribosomes, growth rate, cell mass, protein synthesis speed and metabolic flux (1,30,42).

Conclusion

In summary, we investigated the kinetics of prokaryotic protein synthesis at the ribosomal level and coupled it to cell growth. Translation accuracy couples with the protein synthesis speed at different Mg2+. The speed – accuracy relation of peptide bond formation sheds light on the mechanism of growth rate maximization via tuning translational accuracy and partitioning proteomic content. During the adaptation to the improvement in nutrient quality, cells can tune accuracy to optimize growth rate by regulating Mg2+. Our results thus provide a theoretical framework to understand the mechanism of adaptive response of bacterial cell growth for optimal fitness.

Author contributions

H.S. conceived and supervised the work. L.Q.L. and K.Z. performed research. All authors wrote the manuscript.

Acknowledgments

The authors are grateful to Magnus Johansson, Måns Ehrenberg, Rudy Marcus, Daniela Rhodes, Ada Yonath, and Venki Ramakrishnan for enlightening discussions. H.B.S. is grateful to Weicheng Su and Kerson Huang for enlightening advice at the early stage of this work. We thank Zhihao Liu, Zhouyi He, Mengqi Jia, Chiming Kan, and Shanqi Yap for technique assistance. The authors would like to express their gratitude to the editor Jianhua Xing and anonymous reviewers for their careful reading of our manuscript as well as their insightful comments and suggestions. This work is supported in part by the Southern Marine Science and Engineering Guangdong Laboratory (Guangzhou) (SMSEGL20SC01) and HKUST grant (R9418).

Declaration of interests

The authors declare no competing financial interests.

Notes

Editor: Jianhua Xing.

Footnotes

Luan Quang Le and Kaicheng Zhu contributed equally to this work.

Supporting material can be found online at https://doi.org/10.1016/j.bpj.2022.12.028.

Supporting citations

References (64,65,66,67,68,69,70,71,72,73,74,75,76,77,78) appear in the supporting material.

Supporting material

Document S1. Figures S1–S13 and Tables S1–S6:

Document S2. Article plus supporting material:

References

1. Bremer H., Dennis P. Modulation of chemical composition and other parameters of the cell at different exponential growth rates. EcoSal Plus. 2008;3 doi: 10.1128/ecosal.5.2.3. [PubMed] [CrossRef] [Google Scholar]
2. Dalbow D.G., Young R. Synthesis time of β-galactosidase in Escherichia coli B/r as a function of growth rate. Biochem. J. 1975;150:13–20. [PMC free article] [PubMed] [Google Scholar]
3. Forchhammer J., Lindahl L. Growth rate of polypeptide chains as a function of the cell growth rate in a mutant of Escherichia coli 15. J. Mol. Biol. 1971;55:563–568. [PubMed] [Google Scholar]
4. Reynolds N.M., Lazazzera B.A., Ibba M. Cellular mechanisms that control mistranslation. Nat. Rev. Microbiol. 2010;8:849–856. [PubMed] [Google Scholar]
5. Parker J. Errors and alternatives in reading the universal genetic code. Microbiol. Rev. 1989;53:273–298. [PMC free article] [PubMed] [Google Scholar]
6. Drummond D.A., Wilke C.O. The evolutionary consequences of erroneous protein synthesis. Nat. Rev. Genet. 2009;10:715–724. [PMC free article] [PubMed] [Google Scholar]
7. Bruggeman F.J., Planqué R., et al.Teusink B. Searching for principles of microbial physiology. FEMS Microbiol. Rev. 2020;44:821–844. [PMC free article] [PubMed] [Google Scholar]
8. Thompson R.C., Karim A.M. The accuracy of protein biosynthesis is limited by its speed: high fidelity selection by ribosomes of aminoacyl-tRNA ternary complexes containing GTP[gamma S] Proc. Natl. Acad. Sci. USA. 1982;79:4922–4926. [PMC free article] [PubMed] [Google Scholar]
9. Lovmar M., Ehrenberg M. Rate, accuracy and cost of ribosomes in bacterial cells. Biochimie. 2006;88:951–961. [PubMed] [Google Scholar]
10. Rodnina M.V. Quality Control of mRNA Decoding on the Bacterial Ribosome. Vol. 86. Academic Press; 2012. pp. 95–128. [PubMed] [Google Scholar]
11. Ibarra R.U., Edwards J.S., Palsson B.O. Escherichia coli K-12 undergoes adaptive evolution to achieve in silico predicted optimal growth. Nature. 2002;420:186–189. [PubMed] [Google Scholar]
12. Klumpp S., Scott M., et al.Hwa T. Molecular crowding limits translation and cell growth. Proc. Natl. Acad. Sci. USA. 2013;110:16754–16759. [PMC free article] [PubMed] [Google Scholar]
13. Tubulekas I., Hughes D. Suppression of rpsL phenotypes by tuf mutations reveals a unique relationship between translation elongation and growth rate. Mol. Microbiol. 1993;7:275–284. [PubMed] [Google Scholar]
14. Ehrenberg M., Kurland C.G. Costs of accuracy determined by a maximal growth rate constraint. Q. Rev. Biophys. 1984;17:45–82. [PubMed] [Google Scholar]
15. Zhang J., Ieong K.-W., et al.Ehrenberg M. Accuracy of initial codon selection by aminoacyl-tRNAs on the mRNA-programmed bacterial ribosome. Proc. Natl. Acad. Sci. USA. 2015;112:9602–9607. [PMC free article] [PubMed] [Google Scholar]
16. Johansson M., Zhang J., Ehrenberg M. Genetic code translation displays a linear trade-off between efficiency and accuracy of tRNA selection. Proc. Natl. Acad. Sci. USA. 2012;109:131–136. [PMC free article] [PubMed] [Google Scholar]
17. Rodnina M.V., Wintermeyer W. Fidelity of aminoacyl-tRNA selection on the ribosome: kinetic and structural mechanisms. Annu. Rev. Biochem. 2001;70:415–435. [PubMed] [Google Scholar]
18. Hopfield J.J. Kinetic proofreading: a new mechanism for reducing errors in biosynthetic processes requiring high specificity. Proc. Natl. Acad. Sci. USA. 1974;71:4135–4139. [PMC free article] [PubMed] [Google Scholar]
19. Gromadski K.B., Rodnina M.V. Kinetic determinants of high-fidelity tRNA discrimination on the ribosome. Mol. Cell. 2004;13:191–200. [PubMed] [Google Scholar]
20. Gromadski K.B., Daviter T., Rodnina M.V. A Uniform response to mismatches in codon-anticodon complexes ensures ribosomal fidelity. Mol. Cell. 2006;21:369–377. [PubMed] [Google Scholar]
21. Wang Q., Su H. A tale of water molecules in the ribosomal peptidyl transferase reaction. Biochemistry. 2022;61:2241–2247. [PubMed] [Google Scholar]
22. McClory S.P., Leisring J.M., et al.Fredrick K. Missense suppressor mutations in 16S rRNA reveal the importance of helices h8 and h14 in aminoacyl-tRNA selection. RNA. 2010;16:1925–1934. [PMC free article] [PubMed] [Google Scholar]
23. Ruusala T., Ehrenberg M., Kurland C.G. Is there proofreading during polypeptide synthesis? EMBO J. 1982;1:741–745. [PMC free article] [PubMed] [Google Scholar]
24. Zaher H.S., Green R. Hyperaccurate and error-prone ribosomes exploit distinct mechanisms during tRNA selection. Mol. Cell. 2010;39:110–120. [PMC free article] [PubMed] [Google Scholar]
25. Zhang J., Pavlov M.Y., Ehrenberg M. Accuracy of genetic code translation and its orthogonal corruption by aminoglycosides and Mg2+ ions. Nucleic Acids Res. 2018;46:1362–1374. [PMC free article] [PubMed] [Google Scholar]
26. Ieong K.-W., Uzun l., et al.Ehrenberg M. Two proofreading steps amplify the accuracy of genetic code translation. Proc. Natl. Acad. Sci. USA. 2016;113:13744–13749. [PMC free article] [PubMed] [Google Scholar]
27. Fluitt A., Pienaar E., Viljoen H. Ribosome kinetics and aa-tRNA competition determine rate and fidelity of peptide synthesis. Comput. Biol. Chem. 2007;31:335–346. [PMC free article] [PubMed] [Google Scholar]
28. Schaechter M., MaalØe O., Kjeldgaard N.O. Dependency on medium and temperature of cell size and chemical composition during balanced growth of Salmonella typhimurium. J. Gen. Microbiol. 1958;19:592–606. [PubMed] [Google Scholar]
29. Scott M., Gunderson C.W., et al.Hwa T. Interdependence of cell growth and gene expression: origins and consequences. Science. 2010;330:1099–1102. [PubMed] [Google Scholar]
30. Zheng H., Bai Y., et al.Liu C. General quantitative relations linking cell growth and the cell cycle in Escherichia coli. Nat. Microbiol. 2020;5:995–1001. [PMC free article] [PubMed] [Google Scholar]
31. Bremer H., Dennis P.P. ASM Press; 2009. Escherichia coli and Salmonella: Cellular and Molecular Biology. [Google Scholar]
32. Scott M., Klumpp S., et al.Hwa T. Emergence of robust growth laws from optimal regulation of ribosome synthesis. Mol. Syst. Biol. 2014;10:747. [PMC free article] [PubMed] [Google Scholar]
33. Rosset R., Gorini L. A ribosomal ambiguity mutation. J. Mol. Biol. 1969;39:95–112. [PubMed] [Google Scholar]
34. Hughes D. Error-prone EF-Tu reduces in vivo enzyme activity and cellular growth rate. Mol. Microbiol. 1991;5:623–630. [PubMed] [Google Scholar]
35. Akanuma G. Diverse relationships between metal ions and the ribosome. Biosci. Biotechnol. Biochem. 2021;85:1582–1593. [PubMed] [Google Scholar]
36. Nierhaus K.H. Mg2+, K+, and The ribosome. J. Bacteriol. 2014;196:3817–3819. [PMC free article] [PubMed] [Google Scholar]
37. Alatossava T., Jütte H., et al.Kellenberger E. Manipulation of intracellular magnesium content in polymyxin B nonapeptide-sensitized Escherichia coli by ionophore A23187. J. Bacteriol. 1985;162:413–419. [PMC free article] [PubMed] [Google Scholar]
38. Pech M., Karim Z., et al.Nierhaus K.H. Elongation factor 4 (EF4/LepA) accelerates protein synthesis at increased Mg2+ concentrations. Proc. Natl. Acad. Sci. USA. 2011;108:3199–3203. [PMC free article] [PubMed] [Google Scholar]
39. Nepal S., Kumar P. Growth, cell division, and gene expression of Escherichia coli at elevated concentrations of magnesium sulfate: implications for habitability of europa and mars. Microorganisms. 2020;8:637. [PMC free article] [PubMed] [Google Scholar]
40. Johansson M., Lovmar M., Ehrenberg M. Rate and accuracy of bacterial protein synthesis revisited. Curr. Opin. Microbiol. 2008;11:141–147. [PubMed] [Google Scholar]
41. Zhang J., Ieong K.-W., et al.Ehrenberg M. Proofreading neutralizes potential error hotspots in genetic code translation by transfer RNAs. RNA. 2016;22:896–904. [PMC free article] [PubMed] [Google Scholar]
42. Dai X., Zhu M., et al.Hwa T. Reduction of translating ribosomes enables Escherichia coli to maintain elongation rates during slow growth. Nat. Microbiol. 2016;2:16231. [PMC free article] [PubMed] [Google Scholar]
43. DeLong J.P., Okie J.G., et al.Brown J.H. Shifts in metabolic scaling, production, and efficiency across major evolutionary transitions of life. Proc. Natl. Acad. Sci. USA. 2010;107:12941–12945. [PMC free article] [PubMed] [Google Scholar]
44. Stouthamer A.H. A theoretical study on the amount of ATP required for synthesis of microbial cell material. Antonie Leeuwenhoek. 1973;39:545–565. [PubMed] [Google Scholar]
45. Mori M., Schink S., et al.Hwa T. Quantifying the benefit of a proteome reserve in fluctuating environments. Nat. Commun. 2017;8:1225. [PMC free article] [PubMed] [Google Scholar]
46. Basan M., Zhu M., et al.Hwa T. Inflating bacterial cells by increased protein synthesis. Mol. Syst. Biol. 2015;11:836. [PMC free article] [PubMed] [Google Scholar]
47. Elf J., Berg O.G., Ehrenberg M. Comparison of repressor and transcriptional attenuator systems for control of amino acid biosynthetic operons. J. Mol. Biol. 2001;313:941–954. [PubMed] [Google Scholar]
48. Kempes C.P., Dutkiewicz S., Follows M.J. Growth, metabolic partitioning, and the size of microorganisms. Proc. Natl. Acad. Sci. USA. 2012;109:495–500. [PMC free article] [PubMed] [Google Scholar]
49. Banerjee K., Kolomeisky A.B., Igoshin O.A. Elucidating interplay of speed and accuracy in biological error correction. Proc. Natl. Acad. Sci. USA. 2017;114:5183–5188. [PMC free article] [PubMed] [Google Scholar]
50. Chiuchiú D., Tu Y., Pigolotti S. Error-speed correlations in biopolymer synthesis. Phys. Rev. Lett. 2019;123:038101. [PMC free article] [PubMed] [Google Scholar]
51. Schreier M.H., Staehelin T. Initiation of mammalian protein synthesis: the importance of ribosome and initiation factor quality for the efficiency of in vitro systems. J. Mol. Biol. 1973;73:329–349. [PubMed] [Google Scholar]
52. Manickam N., Nag N., Abbasi A., Patel K., Farabaugh P.J. Studies of translational misreading in vivo show that the ribosome very efficiently discriminates against most potential errors. RNA. 2014;20:9–15. [PMC free article] [PubMed] [Google Scholar]
53. Borg A., Ehrenberg M. Determinants of the rate of mRNA translocation in bacterial protein synthesis. J. Mol. Biol. 2015;427:1835–1847. [PubMed] [Google Scholar]
54. Lee D.-y. D., Galera-Laporta L., et al.Süel G.M. Magnesium flux modulates ribosomes to increase bacterial survival. Cell. 2019;177:352–360.e13. [PMC free article] [PubMed] [Google Scholar]
55. Wendel B.M., Pi H., et al.Helmann J.D. A central role for magnesium homeostasis during adaptation to osmotic stress. mBio. 2022;13 e0009222–22. [PMC free article] [PubMed] [Google Scholar]
56. Hattori M., Iwase N., et al.Nureki O. Mg2+-dependent gating of bacterial MgtE channel underlies Mg2+ homeostasis. EMBO J. 2009;28:3602–3612. [PMC free article] [PubMed] [Google Scholar]
57. Pontes M.H., Yeom J., Groisman E.A. Reducing ribosome biosynthesis promotes translation during low Mg2+stress. Mol. Cell. 2016;64:480–492. [PMC free article] [PubMed] [Google Scholar]
58. García Véscovi E., Soncini F.C., Groisman E.A. Mg2+As an extracellular signal: environmental regulation of Salmonella virulence. Cell. 1996;84:165–174. [PubMed] [Google Scholar]
59. Lusk J.E., Williams R.J., Kennedy E.P. Magnesium and the growth of Escherichia coli. J. Biol. Chem. 1968;243:2618–2624. [PubMed] [Google Scholar]
60. Stoebel D.M., Dean A.M., Dykhuizen D.E. The cost of expression of Escherichia coli lac operon proteins is in the process, not in the products. Genetics. 2008;178:1653–1660. [PMC free article] [PubMed] [Google Scholar]
61. Kurland C.G. Translational accuracy and the fitness of bacteria. Annu. Rev. Genet. 1992;26:29–50. [PubMed] [Google Scholar]
62. Bürger R., Willensdorfer M., Nowak M.A. Why are phenotypic mutation rates much higher than genotypic mutation rates? Genetics. 2006;172:197–206. [PMC free article] [PubMed] [Google Scholar]
63. Wohlgemuth I., Garofalo R., et al.Rodnina M.V. Translation error clusters induced by aminoglycoside antibiotics. Nat. Commun. 2021;12:1830. [PMC free article] [PubMed] [Google Scholar]
64. Lee T.-H., Blanchard S.C., et al.Chu S. The role of fluctuations in tRNA selection by the ribosome. Proc. Natl. Acad. Sci. USA. 2007;104:13661–13665. [PMC free article] [PubMed] [Google Scholar]
65. Piperno J.R., Oxender D.L. Amino acid transport systems in Escherichia coli K12. J. Biol. Chem. 1968;243:5914–5920. [PubMed] [Google Scholar]
66. Milo R. What is the total number of protein molecules per cell volume? A call to rethink some published values. Bioessays. 2013;35:1050–1055. [PMC free article] [PubMed] [Google Scholar]
67. Button D.K. Nutrient uptake by microorganisms according to kinetic parameters from theory as related to cytoarchitecture. Microbiol. Mol. Biol. Rev. 1998;62:636–645. [PMC free article] [PubMed] [Google Scholar]
68. Bosdriesz E., Molenaar D., et al.Bruggeman F.J. How fast-growing bacteria robustly tune their ribosome concentration to approximate growth-rate maximization. FEBS J. 2015;282:2029–2044. [PMC free article] [PubMed] [Google Scholar]
69. Milo R., Phillips R. Cell Biology by the Numbers. Garland Science; 2015. How big is the “average” protein? [Google Scholar]
70. Milo R., Jorgensen P., Moran U., Weber G., Springer M. BioNumbers–the database of key numbers in molecular and cell biology. Nucleic Acids Res. 2010;38:D750–D753. [PMC free article] [PubMed] [Google Scholar]
71. Pape T., Wintermeyer W., Rodnina M. Induced fit in initial selection and proofreading of aminoacyl-tRNA on the ribosome. EMBO J. 1999;18:3800–3807. [PMC free article] [PubMed] [Google Scholar]
72. Koch A.L., Levy H.R. Protein turnover in growing cultures of Escherichia coli. J. Biol. Chem. 1955;217:947–957. [PubMed] [Google Scholar]
73. West G.B., Brown J.H., Enquist B.J. A general model for ontogenetic growth. Nature. 2001;413:628–631. [PubMed] [Google Scholar]
74. Rubin A.H., Terasaki M., Sanui H. Major intracellular cations and growth control: correspondence among magnesium content, protein synthesis, and the onset of DNA synthesis in BALB/c3T3 cells. Proc. Natl. Acad. Sci. USA. 1979;76:3917–3921. [PMC free article] [PubMed] [Google Scholar]
75. Basan M., Hui S., et al.Hwa T. Overflow metabolism in Escherichia coli results from efficient proteome allocation. Nature. 2015;528:99–104. [PMC free article] [PubMed] [Google Scholar]
76. Molenaar D., van Berlo R., et al.Teusink B. Shifts in growth strategies reflect tradeoffs in cellular economics. Mol. Syst. Biol. 2009;5:323. [PMC free article] [PubMed] [Google Scholar]
77. Vemuri G.N., Altman E., et al.Eiteman M.A. Overflow metabolism in Escherichia coli during steady-state growth: transcriptional regulation and effect of the redox ratio. Appl. Environ. Microbiol. 2006;72:3653–3661. [PMC free article] [PubMed] [Google Scholar]
78. Beardmore R.E., Gudelj I., et al.Hurst L.D. Metabolic trade-offs and the maintenance of the fittest and the flattest. Nature. 2011;472:342–346. [PubMed] [Google Scholar]

Articles from Biophysical Journal are provided here courtesy of The Biophysical Society