throbber
Molecular Systems Biology 5; Article number 314; doi:10.1038/msb.2009.68
`Citation: Molecular Systems Biology 5:314
`& 2009 EMBO and Macmillan Publishers Limited All rights reserved 1744-4292/09
`www.molecularsystemsbiology.com
`
`Ribosome and transcript copy numbers, polysome
`occupancy and enzyme dynamics in Arabidopsis
`
`Maria Piques, Waltraud X Schulze, Melanie Ho¨ hne, Bjo¨ rn Usadel, Yves Gibon1, Johann Rohwer2 and Mark Stitt*
`
`Max Planck Institute of Molecular Plant Physiology, Am Muehlenberg 1, Potsdam-Golm, Germany
`1 Present address: INRA Bordeaux, University of Bordeaux 1&2, UMR619 Fruit Biology, F-33883 Villenave d’Ornon, France
`2 Permanent address: Triple-J Group for Molecular Cell Physiology, Department of Biochemistry, Stellenbosch University, Private Bag X1, 7602 Matieland, South Africa
`* Corresponding author. Max Planck Institute of Molecular Plant Physiology, Am Muehlenberg 1, Potsdam-Golm 14474, Germany. Tel.: þ 49 331 5678100;
`Fax: þ 49 331 5678101; E-mail: mstitt@mpimp-golm.mpg.de
`
`Received 24.4.09; accepted 21.7.09
`
`Plants are exposed to continual changes in the environment. The daily alternation between light
`and darkness results in massive recurring changes in the carbon budget, and leads to widespread
`changes in transcript levels. These diurnal changes are superimposed on slower changes in the
`environment. Quantitative molecular information about the numbers of ribosomes, of transcripts
`for 35 enzymes in central metabolism and their loading into polysomes is used to estimate
`translation rates in Arabidopsis rosettes, and explore the consequences for important sub-processes
`in plant growth. Translation rates for individual enzyme are compared with their abundance in the
`rosette to predict which enzymes are subject to rapid turnover every day, and which are synthesized
`at rates that would allow only slow adjustments to sustained changes of the environment, or
`resemble those needed to support the observed rate of growth. Global translation rates are used to
`estimate the energy costs of protein synthesis and relate them to the plant carbon budget, in
`particular the rates of starch degradation and respiration at night.
`Molecular Systems Biology 5: 314; published online 13 October 20009; doi:10.1038/msb.2009.68
`Subject Categories: functional genomics; plant biology
`Keywords: Arabidopsis; polysomes; quantitative RT–PCR; ribosome; translation
`
`This is an open-access article distributed under the terms of the Creative Commons Attribution Licence,
`which permits distribution and reproduction in any medium, provided the original author and source are
`credited. Creation of derivative works is permitted but the resulting work may be distributed only under the
`same or similar licence to this one. This licence does not permit commercial exploitation without specific
`permission.
`
`Introduction
`
`Plant growth is driven by photosynthetic assimilation of
`carbon (C). Nutrients like nitrate are absorbed by the roots and
`converted to amino acids in the leaves using light energy, or
`imported C in the roots. Sucrose and amino acids are
`transported to the shoot and root apex to support growth of
`more leaves and roots. Plants are unavoidably exposed to
`changes in the environment. One of the most striking changes
`is the daily alternation of light and darkness, which leads to
`an extreme and repeated alternation between two states,
`occurs every day in the natural environment, and can be
`precisely simulated in laboratory experiments. It results in a
`large positive balance of energy and C in the light period, and
`a deficit in the dark period. This is buffered by storing some of
`the newly fixed C as starch, and remobilizing it to support
`metabolism and growth during night (Geiger et al, 2000;
`Smith and Stitt, 2007). However, we do not know how
`energetically expensive processes like protein synthesis are
`regulated during these marked diurnal changes in the plant’s
`energy budget.
`
`Plants are also exposed to slower changes that occur in a
`time range of days or weeks, as a result of changing weather
`patterns and seasonal changes. A very large portion of the total
`leaf protein is invested in a single metabolic process. Owing to
`its very low rate of catalysis (Kcat¼3 s
`1), RubisCO represents
`over 30% of the total protein in a leaf (Farquhar et al, 2001; Zhu
`et al, 2007). Large amounts of protein are also invested in the
`synthesis of chlorophyll-binding proteins and other enzymes
`involved in the Calvin cycle and pathways for carbohydrate
`and amino acid synthesis. This raises the question about how
`plants integrate their response to the environment over a wide
`range of time spans to generate appropriate levels of proteins
`for photosynthesis and growth.
`their
`Thousands of genes undergo diurnal changes of
`transcript level (Bla¨sing et al, 2005) driven by the circadian
`clock, light and sugars (Usadel et al, 2008). This includes many
`transcripts that encode enzymes in central metabolism. Gibon
`et al (2004b) developed a robotized platform to profile the
`maximum activities of over 20 enzymes, most of which show
`rather small diurnal changes (Smith et al, 2004; Gibon et al,
`2004b). Although transcripts respond within hours, changes of
`
`& 2009 EMBO and Macmillan Publishers Limited
`
`Molecular Systems Biology 2009 1
`
`Regeneron Exhibit 2017
`Page 01 of 17
`
`

`

`Quantitative polysome analysis in Arabidopsis
`M Piques et al
`
`enzyme abundances require days to adjust when plants are
`transferred to continuous darkness (Gibon et al, 2004b, 2006).
`Gibon et al (2004b) hypothesized that the rate of translation is
`so slow that several days are required to produce a major
`change in protein abundance. As a result, the rapid transient
`changes of transcripts would be integrated over a longer period
`of time to set the levels of enzymes and other proteins. This
`would buffer the enzymatic capacities in central metabolism
`against recurring changes caused by the light–dark cycle,
`while allowing them to adjust to sustained changes in the
`surroundings.
`Protein synthesis occurs by the recruitment of transcripts to
`ribosomes, to form polysomes. Its synthesis represents a major
`portion of the total ATP consumption in animal and plant cells
`(Hachiya et al, 2007; Pace and Manahan, 2007; Proud, 2007).
`Energy is also required for the synthesis of amino acids. The
`conversion of nitrate to amino acids requires about five ATP
`molecules per amino acid (Penning de Vries, 1975; Hachiya
`et al, 2007). The synthesis of ribosomes requires energy, and
`diverts resources from other cellular components. Eukaryotic
`ribosomes typically contain one molecule of each of the four
`different ribosomal RNA (rRNA) species and one molecule
`of ca. 80 different ribosomal proteins (Perry, 2007). Ribosomal
`RNA and proteins represent 480 and 30–50% of the total
`RNA and protein, respectively, in a growing yeast cell (Warner,
`1999; Perry, 2007). There is selective pressure to achieve a
`parsimonious use of the translational machinery (Beilharz and
`Preiss, 2007; Lackner et al, 2007). In budding yeast, up to 85%
`of the ribosomes are present in the polysomes. The ribosome
`density in polysomes is about one-fifth of the theoretical
`maximum (Arava et al, 2003; MacKay et al, 2004), which is
`consistent with the view that translation is generally regulated
`by the rate of initiation. A twofold decrease of transcript levels
`for some ribosomal proteins leads to a ribosome deficit and a
`minute growth reduction phenotype in Drosophila, indicating
`that the overall ribosomal number limits protein synthesis and
`growth (Perry, 2007).
`The rate of translation of a given transcript species depends
`on transcript abundance, the proportion present in polysomes
`(often termed ‘ribosome occupancy’), the number of ribo-
`somes present on the transcript (often termed ‘ribosome
`density’), and their speed of progression along the transcript
`(Arava et al, 2003; Beilharz and Preiss, 2004; Beyer et al, 2004;
`Brockmann et al, 2007). On average, about 70% of a given
`transcript species is occupied by ribosomes. Similar percen-
`tages of ribosomes (60%) and transcripts (59–82%) are loaded
`into polysomes in plants (Kawaguchi et al, 2004; Kawaguchi
`and Bailey-Serres, 2005).
`In the following article, we describe the methods that allow
`quantitative analysis of ribosome and transcript concentra-
`tions, and polysome composition in Arabidopsis rosettes.
`These data allow us to predict translation rates, both globally,
`and for individual enzymes in central metabolism. We explore
`the consequences of these molecular events for important sub-
`processes in plant growth. First, we compare the rates of
`synthesis with the protein abundance to predict which
`enzymes are likely to be subject to rapid turnover. Second,
`we use these molecular data to estimate the costs of protein
`synthesis and relate to the C budget, in particular the rates of
`starch degradation and respiration at night.
`
`Results
`
`Experimental strategy
`
`Figure 1 outlines our experimental strategy. Ribosome copy
`numbers are measured using quantitative real time RT–PCR
`(qRT–PCR) for rRNA, and polysome fractionation was used to
`estimate the proportion actively involved in translation. This
`information is used to estimate the overall rate of protein
`synthesis. qRT–PCR is combined with polysome fractionation
`to estimate the copy number of transcripts in polysomes,
`including 84 that encode enzymes involved in central plant
`metabolism. This information is used to estimate the rates of
`translation of individual transcripts. In parallel, quantitative
`proteomics and robotized measurements of maximum enzyme
`activities are used to provide two independent estimates of the
`amounts of these enzymes in the leaves. By comparing the
`estimated rates of synthesis with the measured abundance of
`total and individual proteins, and the rate of growth, it is
`possible to predict whether the global rate of protein synthesis
`and the rates of synthesis of individual proteins are of the same
`order as that requited for growth, or whether proteins are
`subject to rapid turnover.
`
`Polysome fractionation
`
`In preliminary experiments, we collected whole rosettes from
`5-week-old wild-type Arabidopsis growing in a 12-h light–dark
`cycle at 6 times during the diurnal cycle, fractionated the
`material by centrifugation, and collected three fractions: the
`
`Figure 1 Quantitative analysis of translation in Arabidopsis rosette leaves.
`
`2 Molecular Systems Biology 2009
`
`& 2009 EMBO and Macmillan Publishers Limited
`
`Regeneron Exhibit 2017
`Page 02 of 17
`
`

`

`Table I Ribosome content in the cytosol, plastids and mitochondria
`1 FW)Ribosome content (mol g
`
`
`
`Cytosol
`Plastid
`Mitochondrion
`
`Dark period
`7.62E11±1.56E11
`2.64E11±6.66E12
`2.25E12±4.42E13
`
`Light period
`7.27E11±1.79E11
`2.57E11±7.42E12
`2.21E12±3.96E13
`
`The results are represented as mean±s.d. of three biological samples.
`Estimation was done by qRT–PCR of SSU rRNA subunits of the cytosolic,
`plastidic and mitochondrial ribosomes.
`
`non-polysomal fraction (NPS) at the top of the gradient,
`the small polysomal fraction (SPS) with an estimated 2–4
`ribosomes per polysome, and the large polysomal fraction
`(LPS) with an estimated X5 ribosomes per polysome. The
`distribution of total RNA in polysome gradients was monitored
`by measuring A254 (Supplementary Figure 1). The proportion
`in the LPS decreased by about twofold during the night, and
`recovered within 2 h in the next light period (Supplementary
`Figure 2). In subsequent experiments, plants were collected at
`the end of the night (‘dark’) period and after 2 h of illumination
`(‘light’), corresponding to the largest changes in polysome
`abundance.
`
`Quantification of ribosomes
`
`A254 does not distinguish between transcript RNA and rRNA,
`and does not distinguish between cytosolic, plastid and
`mitochondrial rRNA. Two complementary approaches were
`taken to quantify cytosolic and plastid ribosomes.
`The first approach used qRT–PCR to determine the
`concentrations of 18S, 16S and 18S-like rRNAs, corresponding
`to the rRNA in the small subunit of the cytosolic, plastid
`and mitochondrial ribosomes, respectively. To allow precise
`quantification, the qRT–PCR data were normalized on four
`artificial control RNAs, which were added at a known
`concentration before purification of
`the rRNA from the
`gradient
`fractions. The total estimated concentration of
`1 fresh weight (FW) (see,
`ribosomes is about 0.10 nmol g
`Table I and Calculations and assumptions section). Cytosolic
`ribosomes were threefold more abundant than plastid ribo-
`somes, and 30-fold more abundant
`than mitochondrial
`ribosomes. This is in agreement with earlier reports that
`26–36% of the total ribosomes in leaves are present in plastids
`(Dyer et al, 1971). The distribution of different rRNA species in
`the polysome gradient suggests that about twofold more
`ribosomes were present in the polysomes in the light period
`than in the dark period (Figure 2A and C). The proportion of
`cytosolic rRNA in the LPS fraction increased from 26 to 50%,
`whereas the proportion in SPS decreased slightly (from 27 to
`20%), as expected in the case if the polysome population is
`shifting towards a higher number of ribosomes per transcript,
`and free cytosolic ribosomes decreased from 47 to 30%. A
`qualitatively similar picture was found for plastid rRNA,
`except that a smaller proportion was found in polysomes
`(Figure 2B). This conclusion is supported by the distribution of
`ribosomal proteins (Figure 2D). The distribution of mitochon-
`drial rRNA was not investigated.
`
`Quantitative polysome analysis in Arabidopsis
`M Piques et al
`
`Figure 2 Distribution of ribosomes in different polysomal fractions in the dark
`and the light periods. (A, B) Ribosome number in each fraction was calculated
`by determining the amount of the SSU rRNA for cytosolic and plastid ribosomes
`by qRT–PCR, assuming each rRNA copy corresponds to one ribosome. (C, D)
`Ribosomal protein abundance in each fraction, calculated by normalizing the
`summed emPAI values for ribosomal proteins on total protein in the fraction.
`
`The second approach used relative quantitative proteomics
`based on emPAI values (Ishihama et al, 2005) to estimate the
`abundance of ribosomal proteins in the polysome gradient
`fractions (Figure 2C and D). Briefly, the number of identified
`peptides per protein are corrected for the number of possible
`peptides from that protein, and taken as quantitative measure
`of the protein abundance (see Calculations and assumptions
`section). This approach confirmed that about twofold more
`ribosomes are present in polysomes in the light period than in
`the dark period, and that a larger proportion of cytosolic
`ribosomes than plastid ribosomes are present in polysomes.
`The measurements of protein abundance indicate a higher
`proportion of ribosomes in polysomes than the measurements
`of rRNA for cytosolic ribosomes. This might be for technical
`reasons. The NPS fraction is a complex matrix with a high
`proportion of proteins with other biological functions and
`emPAI may underestimate ribosomal proteins, especially in
`the light period, when they are strongly depleted in NPS.
`Taken together, both approaches reveal marked changes in
`ribosomal loading into polysomes between the dark period
`and the light period, with 53–58% of the cytosolic ribosomes
`being loaded in the dark period, and 70–90% in the light
`period. The values in the light are similar to those in rapidly
`growing yeast (see Introduction).
`
`Global rates of protein synthesis compared with
`growth requirements
`
`ribosomes and the
`the numbers of
`Information about
`proportion found in polysomes was used to estimate the
`overall rate of protein synthesis. The summed concentration of
`
`& 2009 EMBO and Macmillan Publishers Limited
`
`Molecular Systems Biology 2009 3
`
`Regeneron Exhibit 2017
`Page 03 of 17
`
`

`

`Quantitative polysome analysis in Arabidopsis
`M Piques et al
`
`cytosolic, plastid and mitochondrial ribosomes is about
`1 FW (see Table I). Assuming that a ribosome
`0.1 nmol g
`adds 3 amino acids per s (see Calculations and assumptions
`section), these could catalyze the addition of 26 mmol amino
`1, equivalent to the synthesis of about 3 mg
`acids per g FW day
`1 FW day
`1. The actual rate of protein synthesis is
`protein g
`lower, because only 70% of the ribosomes are in polysomes in
`the light period, and 40% in the dark period (Figure 2),
`resulting in an estimated synthesis rate of 1.8 mg pro-
`
`1 FW day1. Under the conditions used in these experi-
`tein g
`1 FW
`ments, Arabidopsis contains about 15 mg protein g
`(Gibon et al, 2004a, 2009; Hannemann et al, 2009) and grows
`exponentially (see e.g. Tschoep et al, 2009), with a relative
`
`1 FW day1, which is equiva-
`growth rate of 0.15–0.20 g FW g
`1 FW day
`1. The
`lent to the synthesis of 2.2–3.0 mg protein g
`rate of protein synthesis estimated from ribosome copy
`number therefore resembles that required for the observed
`rate of growth.
`
`Ribosomal occupancy of transcripts encoding
`enzymes in central metabolism
`
`We next investigated ribosomal occupancy of transcripts for 98
`genes (Supplementary Table I) including the major members
`of the gene families for 35 enzymes of central metabolism (84
`transcripts). We also included AtCAB1 and AtCAB2 as
`representatives of photosynthetic and circadian-regulated
`genes whose expressions peak at the beginning of the day
`(Ernst et al, 1990), three circadian-regulated genes whose
`expressions peak at the end of the day (GER3, CAT3, GRP7) and
`9 transcripts for ‘house-keeping’ genes that are frequently used
`for normalization of transcriptional analyses (Czechowski
`et al, 2005). They are all nuclear encoded, except for the
`plastid-encoded RubisCO large subunit (RBCL).
`We used qRT–PCR to investigate the distribution of these 98
`transcripts in the polysome gradients. To allow precise
`quantification, the qRT–PCR data were normalized on four
`artificial control RNAs that were added at a known concentra-
`tion before purification of RNA in the gradient fractions. This
`allowed copy numbers of each transcript species to be
`determined per fraction. In addition to being necessary for
`subsequent calculations (see below),
`this bypassed the
`problems associated with normalization of transcript levels
`between fractions in polysome gradients. Transcript levels are
`usually normalized to total RNA. However, the relative levels
`of transcript and rRNA probably change across a polysome
`gradient. ‘House-keeping’ genes are often used for normal-
`ization of qRT–PCR data between different organs or environ-
`ment treatments, but cannot be used for this purpose in
`polysome fractions, because it is not known whether they are
`subject to translational regulation. The proportion of transcript
`in each fraction of the polysome gradient can be easily
`calculated using qRT–PCR data in combination with internal
`standards, by simply comparing the numbers of transcripts
`against
`internal standard as a control across the whole
`gradient. The transcript concentrations in the rosette (un-
`fractionated RNA) and polysome fractions, and the ribosomal
`occupancy (i.e. the proportion of a transcript in the SPS and
`LPS fractions) are provided in Supplementary Table II, along
`
`Figure 3 Scatter plot comparing the ribosome occupancy of 98 transcripts in
`the night and in the light period. Ribosome occupancy was calculated as
`(SPSþ LPS)/(NPS þ SPSþ LPS). Green circles represent photosynthetic
`proteins, green filled circles indicate RBCS gene family and RBCL, and red
`circles indicate genes that are classified as ‘house-keeping’
`in expression
`studies. The plot is generated using data provided in Supplementary Table II.
`
`with the information about the length of the coding sequence,
`and length and molecular weight of the encoded peptide.
`Ribosomal occupancy of the 98 transcripts analyzed, varied
`between 40–95% in the dark period, and 50–90% in the light
`period (Figure 3). Most transcripts show a marked increase in
`occupancy between the dark and the light periods, with an
`increase of between 5 and 55% in absolute terms, and between
`10 and 100% relative to the value in the dark period. ‘House-
`keeping’ genes showed a similar response to other transcripts.
`RBCL, RBCS-1B and RBCS-2B did not show any increase in
`ribosomal occupancy in the light (see below for further
`discussion). The increase in ribosomal occupancy in the light
`is smaller than the twofold increase of ribosome in polysomes
`(Figure 2), indicating that the average number of ribosomes
`per transcript probably increases in the light period.
`Ribosomal occupancy was weakly but significantly depen-
`dent on transcript concentration, with a Pearson’s R2 value of
`0.065 (P¼0.011) in the dark period, and 0.102 (P¼0.001) in the
`light period (Figure 4). Transcript concentrations varied by
`three orders of magnitude. There was a large range of
`ribosomal occupancy (from 40 to 480%) for transcripts with
`the same concentration. Some of this was due to the effect of
`light, but there was still a large range when transcripts are
`considered in one condition. This indicates that ribosomal
`occupancy depends more on individual features of transcripts
`than their concentrations.
`For almost all transcripts except those encoding RubisCO
`subunits, there are 10–100 times lesser transcripts in the SPS
`fraction than in the LPS fraction. This is shown by the low log10
`ratio of SPS/LPS in Figure 5. The proportion of transcript in
`SPS compared with LPS decreases as the proportion in the NPS
`fraction decreases, resulting in a near-linear relation in this
`semi-log plot. This empirical relationship indicates that
`
`4 Molecular Systems Biology 2009
`
`& 2009 EMBO and Macmillan Publishers Limited
`
`Regeneron Exhibit 2017
`Page 04 of 17
`
`

`

`Quantitative polysome analysis in Arabidopsis
`M Piques et al
`
`density of ribosomes per transcript, resulting in a decrease of
`the proportion found in the SPS compared with the LPS
`fraction. However, the amount of transcript in the NPS fraction
`is higher than would be expected from a simple binomial
`distribution (data not shown); indicating that the probability
`that ribosomes are recruited to a free transcript is lower than
`for a transcript that already has bound ribosomes.
`The five solid symbols in Figures 3–5 depict the response of
`RBCS-1A, RBCS-1B, RBCS-2B, RBCS-3B (the small nuclear-
`encoded subunit of RubisCO) and RBCL (the large plastid-
`encoded subunit of RubisCO). These five transcripts deviate
`strongly from the general relationship between NPS and SPS/
`LPS. These five transcripts show a high occupancy in the
`dark as well as the light period, and a large proportion of the
`transcript is present in the SPS fraction (18–39% and 9–25% in
`the dark period and light period, respectively) (Figure 5). This
`could indicate that RBCL and RBCS transcripts are subject to
`complex translational control (see Discussion section).
`
`Estimation of translation rates
`
`The rate of synthesis of the proteins encoded by these 98
`transcripts was estimated from the transcript abundance in the
`SPS and LPS fractions, multiplied by the ribosome density per
`translating transcript (see Calculations and assumptions
`section, and Supplementary Table II). The calculation assumes
`an elongation rate of 3 amino acids per ribosome per s, an
`average of 3 ribosomes per transcript in the SPS fraction, and a
`ribosome density of 6.6 ribosomes per kb coding sequence,
`(Brandt et al, 2009) in the LPS fraction.
`The estimated rates of protein synthesis (mol pro-
`1) ranged from 2.5E15 to 2.9E09 in the dark
`1 FW h
`tein g
`period, and 6.5E15 to 4.3E10 in the light period (Table II
`and Supplementary Table II). Among the enzymes involved in
`primary metabolism, RubisCO is the most rapidly synthesized
`enzyme, reflecting the high abundance of this enzyme in
`leaves (see Introduction section). Other rapidly synthesized
`enzymes include several Calvin cycle enzymes (e.g. aldolase,
`NADP–GAPDH, PGK and TPI), enzymes involved in nitrogen
`assimilation (e.g. AlaAT, NR and GS), and NAD–GAPDH and
`NAD–MDH. The relatively high rates of synthesis of PEP
`carboxylase, NADP–IDH, aconitase and PK compared with
`other glycolytic enzymes may reflect the fact that
`these
`enzymes are required to synthesize 2-oxoglutarate, which acts
`as the C acceptor during nitrate and ammonium assimilation.
`The relatively high rate of synthesis of glycerate kinase may be
`related to the fact that in leaves this enzyme is required for
`photorespiration. Fluxes through this pathway are roughly
`15–20% of those through photosynthesis (Zhu et al, 2007),
`and much higher than in respiratory metabolism. Most
`enzymes showed an estimated 50–100% increase in the rate
`of synthesis in the light period compared to the dark period. To
`further interpret the biological significance of these rates of
`synthesis, we compared them with the estimated amount of
`each enzyme in the rosette.
`
`Estimation of protein abundance
`
`Protein abundance of metabolic enzymes in rosette leaves was
`estimated by two independent methods. In one approach,
`
`Figure 4 Scatter plot for transcript abundance versus ribosome occupancy in
`the night and the light period. Ribosomal occupancy was calculated as
`(SPSþ LPS)/(NPS þ SPSþ LPS). Blue and orange symbols denote plant
`material collected in the dark and light periods, respectively. Filled symbols
`denote the RBCS gene family (K) and RBCL (’). Ribosomal occupancy is
`weakly, but significantly dependent on transcript concentration in the dark period
`(Pearson’s R2¼0.065, P-value¼0.011) and light period (Pearson’s R2¼0.102,
`P-value¼0.001). The plot
`is generated by using the data provided in
`Supplementary Table II.
`
`Figure 5 Relation between the fraction of transcript in the non-polysomal
`fraction (NPS) and the distribution of transcript between the small (SPS) and
`large polysomal (LPS) fractions. Blue and orange symbols denote plant material
`collected in the dark and light periods, respectively. Filled symbols denote the
`RBCS gene family (K) and RBCL (’). The plot is generated using data
`provided in Supplementary Table II.
`
`initiation and ribosome progression are determined in a
`similar manner for all these transcripts. It is consistent with
`initiation being the limiting step; an increased probability of
`initiation will result in a decreased fraction of the transcript in
`the NPS fraction, and will also result in an increased average
`
`& 2009 EMBO and Macmillan Publishers Limited
`
`Molecular Systems Biology 2009 5
`
`Regeneron Exhibit 2017
`Page 05 of 17
`
`

`

`Quantitative polysome analysis in Arabidopsis
`M Piques et al
`
`Table II Estimated rates of protein synthesis of the different enzymes in
`Arabidopsis rosette in the dark and light periods
`
`Enzyme
`
`Ribulose-1,5-bisphosphate
`carboxylase (RubisCO)
`Fructose-bisphosphate aldolase
`(aldolase)
`NADP–glyceraldehyde 3-
`phosphate dehydrogenase
`(NADP–GAPDH)
`Alanine aminotransferase (AlaAT)
`NAD–glyceraldehyde 3-phosphate
`dehydrogenase (NAD–GAPDH)
`NAD–malate dehydrogenase
`(NAD–MDH)
`Nitrate reductase (NR)
`Glutamine synthetase (GS)
`Phosphoglycerokinase (PGK)
`Triose phosphateisomerase (TPI)
`ADP-glucose pyrophosphorylase
`(AGPase)
`Phosphoenolpyruvate carboxylase
`(PEP carboxylase)
`NADP–isocitrate dehydrogenase
`(NADP–IDH)
`Transketolase (TK)
`Aconitase
`Pyruvate kinase (PK)
`NADP–malate dehydrogenase
`(NADP–MDH)
`Acid invertase (INV)
`Glycerate kinase (GK)
`Glucose-6-phosphate isomerase
`(PGI)
`UDP-glucose pyrophosphorylase
`(UGPase)
`Ferredoxin–glutamate synthase
`(Fd–GOGAT)
`Phosphoglucomutase (PGM)
`PPi-phosphofructokinase (PFP)
`Fructose-1,6-bisphosphatase,
`cytosolic (cytFBPase)
`Sucrose phosphate synthase (SPS)
`Fructokinase (FK)
`Fumarase (FUM)
`Glucose-6-phosphate
`dehydrogenase (G6PDH)
`Aspartate aminotransferase
`(AspAT)
`NAD–isocitrate dehydrogenase
`(NAD–IDH)
`NAD–glutamate dehydrogenase
`(NAD–GDH)
`Glucokinase/hexokinase (HK)
`ATP-phosphofructokinase (PFK)
`Shikimate 5-dehydrogenase
`(Shikimate DH)
`
`Estimated translation rate
`
`1 g1 FW)
`(mol h
`
`Dark period
`
`Light period
`
`6.34E–10
`
`8.60E–10
`
`2.84E–11
`
`5.97E–11
`
`2.94E–11
`
`5.13E–11
`
`2.35E–11
`1.48E–11
`
`3.40E–11
`2.60E–11
`
`1.60E–11
`
`2.03E–11
`
`2.30E–11
`1.41E–11
`1.33E–11
`8.24E–12
`3.54E–12
`
`9.20E–12
`1.42E–11
`1.41E–11
`1.18E–11
`5.94E–12
`
`3.90E–12
`
`4.75E–12
`
`5.11E–12
`
`3.35E–12
`
`4.01E–12
`2.45E–12
`2.27E–12
`2.27E–12
`
`2.29E–12
`1.32E–12
`1.09E–12
`
`3.58E–12
`4.44E–12
`3.07E–12
`3.05E–12
`
`2.05E–12
`1.56E–12
`1.77E–12
`
`9.04E–13
`
`1.63E–12
`
`1.39E–12
`
`1.14E–12
`
`8.23E–13
`7.25E–13
`6.29E–13
`
`7.92E–13
`7.56E–13
`4.92E–13
`7.57E–13
`
`1.44E–12
`1.18E–12
`1.25E–12
`
`9.80E–13
`9.20E–13
`9.92E–13
`6.36E–13
`
`4.96E–13
`
`6.22E–13
`
`4.49E–13
`
`5.69E–13
`
`5.46E–13
`
`3.08E–13
`
`2.40E–13
`3.06E–13
`2.34E–13
`
`4.57E–13
`3.83E–13
`3.62E–13
`
`The raw data and calculations are provided in Supplementary Table II.
`
`protein abundance was estimated from the maximum enzyme
`activity (original data provided in Supplementary Table III),
`corrected by literature values for the specific activity (Supple-
`mentary Table IV). In case of disagreement, the highest specific
`activity was chosen because an underestimate of the specific
`activity can easily occur due to loss of activity or incomplete
`purification. Enzyme activities were measured using a robot-
`ized enzyme determination platform, in which products are
`quantitatively determined in highly sensitive cycling assays
`
`Figure 6 Relation between protein concentrations of metabolic enzymes
`calculated from specific enzyme activities or from the emPAI protein abundance
`index determined by mass spectrometric analysis. Filled symbol denotes
`RubisCO, which is not included in the Pearson’s regression analysis. Protein
`abundances estimated using the two methods are highly significantly correlated
`(R2¼0.592, P-value¼0.681E06). The plot is generated by using the data
`provided in Supplementary Tables V and VII.
`
`that allow high dilution ratios and minimize the interference
`form other components in the extracts (Gibon et al, 2004b). For
`all assays, it was checked that the substrate concentration was
`saturating, and that the activity was linear with the amount of
`extract added. The results and calculations are summarized in
`Supplementary Table V.
`In the second approach, relative protein amounts were
`estimated using the emPAI index (Supplementary Table VI).
`The emPAI index is calculated from the fraction of the number
`of experimentally identified tryptic peptides out of all
`detectable tryptic peptides within the mass range of the mass
`spectrometer (Ishihama et al, 2005). This method yields
`abundances that are linear over three orders of magnitude in
`1–10 mmol l
`1), although it
`protein concentration (10 nmol l
`does underestimate very highly abundant single proteins
`(Ishihama et al, 2005). In our data set, RubisCO subunits
`summed to 410% of all proteins identified, and are probably
`underestimated (Supplementary Table VI). The experimen-
`1 FW) was
`tally determined total protein content (15 mg g
`used to convert the mol% values to concentrations (Supple-
`mentary Table VI). On a molar basis, proteins with photo-
`synthetic functions contributed 30% of all identified proteins.
`Protein synthesis made up the second largest functional
`protein group (11%). Proteins with functions in amino acid
`metabolism, redox regulation and TCA cycle/organic acid
`transformation contributed with 5–6% each (Supplementary
`Table VI).
`Both approaches are subject to experimental error and
`involve assumptions. Nevertheless, comparison of the result-
`ing estimates for protein abundance (Figure 6 and Table III)
`revealed a highly significant agreement (Pearson’s R2 ¼0.590,
`P¼0.727E06). The slope of the regression (0.73) deviated
`from 1, with a smaller dynamic range for the proteomics-based
`
`6 Molecular Systems Biology 2009
`
`& 2009 EMBO and Macmillan Publishers Limited
`
`Regeneron Exhibit 2017
`Page 06 of 17
`
`

`

`Quantitative polysome analysis in Arabidopsis
`M Piques et al
`
`Table III Comparison of the estimated enzyme abundance in the rosette and the corresponding estimated time to synthesize the entire enzyme in the rosette in days
`(TP)
`
`Enzyme
`
`1 FW)Protein abundance (mol g
`
`
`
`Days to synthesize all the protein in
`the rosette (TP)
`
`Nitrate reductase (NR)
`ATP-phosphofructokinase (PFK)
`Acid Invertase (INV)
`Glucose-6-phosphate dehydrogenase (G6PDH)
`Glucokinase/hexokinase (HK)
`Alanine aminotransferase (AlaAT)
`ADP-glucose pyrophosphorylase (AGPase)
`Pyrophosphate-phosphofructokinase (PFP)
`NADP–glyceraldehyde 3-phosphate dehydrogenase
`(NADP–GAPDH)
`Ribulose-1,5-bisphosphate carboxylase (RuBisCO)
`Glucose-6-phosphate isomerase (PGI)**
`Phosphoenolpyruvate carboxylase (PEP carboxylase)**
`Triose phosphate isomerase (TPI)*
`Glutamine synthetase (GS)
`Pyruvate kinase (PK)*
`Fructose-1,6-bisphosphatase, cytosolic (cytFBPase)*
`Sucrose phosphate synthase (SPS)
`NAD–isocitrate dehydrogenase (NAD–IDH)
`NAD–glyceraldehyde 3-phosphate dehydrogenase
`(NAD–GAPDH)
`NADP–isocitrate dehydrogenase (NADP–IDH)
`Shikimate 5-dehydrogenase (Shik

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket