SBMLBioModels: M - M

M


Mardinoglu2014 - Genome-scale metabolic model (HMR version 2.0) - human hepatocytes (iHepatocytes2322): MODEL1402200003v0.0.1

Mardinoglu2014 - Genome-scale metabolic model (HMR version 2.0) - human hepatocytes (iHepatocytes2322)This model is desc…

Details

Several liver disorders result from perturbations in the metabolism of hepatocytes, and their underlying mechanisms can be outlined through the use of genome-scale metabolic models (GEMs). Here we reconstruct a consensus GEM for hepatocytes, which we call iHepatocytes2322, that extends previous models by including an extensive description of lipid metabolism. We build iHepatocytes2322 using Human Metabolic Reaction 2.0 database and proteomics data in Human Protein Atlas, which experimentally validates the incorporated reactions. The reconstruction process enables improved annotation of the proteomics data using the network centric view of iHepatocytes2322. We then use iHepatocytes2322 to analyse transcriptomics data obtained from patients with non-alcoholic fatty liver disease. We show that blood concentrations of chondroitin and heparan sulphates are suitable for diagnosing non-alcoholic steatohepatitis and for the staging of non-alcoholic fatty liver disease. Furthermore, we observe serine deficiency in patients with NASH and identify PSPH, SHMT1 and BCAT1 as potential therapeutic targets for the treatment of non-alcoholic steatohepatitis. link: http://identifiers.org/pubmed/24419221

Mardinoglu2015 - Curated tissue-specific genome-scale metabolic model - Small intestine: MODEL1509220032v0.0.1

Mardinoglu2015 - Curated tissue-specific genome-scale metabolic model - Small intestineThis model is described in the ar…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Curated tissue-specific genome-scale metabolic network - Colon: MODEL1509220030v0.0.1

Genome-scale metabolic model for mouse colon tissueThis model is described in the article: [The gut microbiota modulate…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Curated tissue-specific genome-scale metabolic network - Adipose: MODEL1509220031v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Adipose tissueThis model is described in the article:…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Curated tissue-specific genome-scale metabolic network - Liver: MODEL1509220029v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - LiverThis model is described in the article: [The gut…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Generic mouse genome-scale metabolic network (MMR): MODEL1509220028v0.0.1

Mardinoglu2015 - Generic mouse genome-scale metabolic network (MMR)This model is described in the article: [The gut mic…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Adrenal gland: MODEL1509220011v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Adrenal glandThis model is described in the article:…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Brain cortex: MODEL1509220013v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Brain cortexThis model is described in the article: […

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Brain medulla: MODEL1509220000v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Brain medullaThis model is described in the article:…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Brown fat: MODEL1509220006v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Brown fatThis model is described in the article: [The…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Cerebellum: MODEL1509220002v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - CerebellumThis model is described in the article: [Th…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Colon: MODEL1509220003v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - ColonThis model is described in the article: [The gut…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Diaphragm: MODEL1509220005v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - DiaphragmThis model is described in the article: [The…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Duodenum: MODEL1509220010v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - DuodenumThis model is described in the article: [The…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Embryonic tissue: MODEL1509220001v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Embryonic tissueThis model is described in the article…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Eye: MODEL1509220004v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - EyeThis model is described in the article: [The gut m…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Heart: MODEL1509220014v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - HeartThis model is described in the article: [The gut…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Ileum: MODEL1509220009v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - IleumThis model is described in the article: [The gut…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Jejunum: MODEL1509220007v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - JejunumThis model is described in the article: [The g…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Kidney cortex: MODEL1509220012v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Kidney cortexThis model is described in the article:…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Kidney medulla: MODEL1509220008v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Kidney medullaThis model is described in the article:…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Liver: MODEL1509220017v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - LiverThis model is described in the article: [The gut…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Lung: MODEL1509220015v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - LungThis model is described in the article: [The gut…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Midbrain: MODEL1509220018v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - MidbrainThis model is described in the article: [The…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Muscle: MODEL1509220016v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - MuscleThis model is described in the article: [The gu…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Olfactory bulb: MODEL1509220025v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Olfactory bulbThis model is described in the article:…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Ovary: MODEL1509220023v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - OvaryThis model is described in the article: [The gut…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Pancreas: MODEL1509220024v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - PancreasThis model is described in the article: [The…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Salivary gland: MODEL1509220022v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Salivary glandThis model is described in the article:…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Spleeen: MODEL1509220020v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - SpleeenThis model is described in the article: [The g…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Stomach: MODEL1509220021v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - StomachThis model is described in the article: [The g…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Thymus: MODEL1509220019v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - ThymusThis model is described in the article: [The gu…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - Uterus: MODEL1509220027v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - UterusThis model is described in the article: [The gu…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - White fat: MODEL1509220026v0.0.1

Mardinoglu2015 - Tissue-specific genome-scale metabolic network - White fatThis model is described in the article: [The…

Details

The gut microbiota has been proposed as an environmental factor that promotes the progression of metabolic diseases. Here, we investigated how the gut microbiota modulates the global metabolic differences in duodenum, jejunum, ileum, colon, liver, and two white adipose tissue depots obtained from conventionally raised (CONV-R) and germ-free (GF) mice using gene expression data and tissue-specific genome-scale metabolic models (GEMs). We created a generic mouse metabolic reaction (MMR) GEM, reconstructed 28 tissue-specific GEMs based on proteomics data, and manually curated GEMs for small intestine, colon, liver, and adipose tissues. We used these functional models to determine the global metabolic differences between CONV-R and GF mice. Based on gene expression data, we found that the gut microbiota affects the host amino acid (AA) metabolism, which leads to modifications in glutathione metabolism. To validate our predictions, we measured the level of AAs and N-acetylated AAs in the hepatic portal vein of CONV-R and GF mice. Finally, we simulated the metabolic differences between the small intestine of the CONV-R and GF mice accounting for the content of the diet and relative gene expression differences. Our analyses revealed that the gut microbiota influences host amino acid and glutathione metabolism in mice. link: http://identifiers.org/pubmed/26475342

Maree2006_DuCa_Type1DiabetesModel: BIOMD0000000381v0.0.1

This a model from the article: Modelling the onset of Type 1 diabetes: can impaired macrophage phagocytosis make the…

Details

A wave of apoptosis (programmed cell death) occurs normally in pancreatic beta-cells of newborn mice. We previously showed that macrophages from non-obese diabetic (NOD) mice become activated more slowly and engulf apoptotic cells at a lower rate than macrophages from control (Balb/c) mice. It has been hypothesized that this low clearance could result in secondary necrosis, escalating inflammation and self-antigen presentation that later triggers autoimmune, Type 1 diabetes (T1D). We here investigate whether this hypothesis could offer a reasonable and parsimonious explanation for onset of T1D in NOD mice. We quantify variants of the Copenhagen model (Freiesleben De Blasio et al. 1999 Diabetes 48, 1677), based on parameters from NOD and Balb/c experimental data. We show that the original Copenhagen model fails to explain observed phenomena within a reasonable range of parameter values, predicting an unrealistic all-or-none disease occurrence for both strains. However, if we take into account that, in general, activated macrophages produce harmful cytokines only when engulfing necrotic (but not apoptotic) cells, then the revised model becomes qualitatively and quantitatively reasonable. Further, we show that known differences between NOD and Balb/c mouse macrophage kinetics are large enough to account for the fact that an apoptotic wave can trigger escalating inflammatory response in NOD, but not Balb/c mice. In Balb/c mice, macrophages clear the apoptotic wave so efficiently, that chronic inflammation is prevented. link: http://identifiers.org/pubmed/16608707

Parameters:

NameDescription
Amax = 2.0E7; f2 = 1.0E-5; W = 4936.39216346718; kc = 1.0; f1 = 1.0E-5; d = 0.5Reaction: Ba = (W+Amax*Cy/(kc+Cy))-(f1*M+f2*Ma+d)*Ba, Rate Law: (W+Amax*Cy/(kc+Cy))-(f1*M+f2*Ma+d)*Ba
f2 = 1.0E-5; f1 = 1.0E-5; d = 0.5Reaction: Bn = d*Ba-(f1*M+f2*Ma)*Bn, Rate Law: d*Ba-(f1*M+f2*Ma)*Bn
b = 0.09; J = 50000.0; f1 = 1.0E-5; e1 = 1.0E-8; c = 0.1; k = 0.4Reaction: M = (((J+(k+b)*Ma)-c*M)-f1*M*Ba)-e1*M*(M+Ma), Rate Law: (((J+(k+b)*Ma)-c*M)-f1*M*Ba)-e1*M*(M+Ma)
delta = 25.0; alpha = 5.0E-9Reaction: Cy = alpha*Bn*Ma-delta*Cy, Rate Law: alpha*Bn*Ma-delta*Cy
f1 = 1.0E-5; k = 0.4; e2 = 1.0E-8Reaction: Ma = (f1*M*Ba-k*Ma)-e2*Ma*(M+Ma), Rate Law: (f1*M*Ba-k*Ma)-e2*Ma*(M+Ma)

States:

NameDescription
M[macrophage]
Ma[macrophage]
CyCy
Bn[pancreatic beta cell]
Ba[pancreatic beta cell]

Marhl2000_CaOscillations: BIOMD0000000039v0.0.1

In order to reproduce the model, the volume of all compartment is set to 1, and the stoichiometry of CaER and CaM has be…

Details

Intracellular calcium oscillations, which are oscillatory changes of cytosolic calcium concentration in response to agonist stimulation, are experimentally well observed in various living cells. Simple calcium oscillations represent the most common pattern and many mathematical models have been published to describe this type of oscillation. On the other hand, relatively few theoretical studies have been proposed to give an explanation of complex intracellular calcium oscillations, such as bursting and chaos. In this paper, we develop a new possible mechanism for complex calcium oscillations based on the interplay between three calcium stores in the cell: the endoplasmic reticulum (ER), mitochondria and cytosolic proteins. The majority ( approximately 80%) of calcium released from the ER is first very quickly sequestered by mitochondria. Afterwards, a much slower release of calcium from the mitochondria serves as the calcium supply for the intermediate calcium exchanges between the ER and the cytosolic proteins causing bursting calcium oscillations. Depending on the permeability of the ER channels and on the kinetic properties of calcium binding to the cytosolic proteins, different patterns of complex calcium oscillations appear. With our model, we are able to explain simple calcium oscillations, bursting and chaos. Chaos is also observed for calcium oscillations in the bursting mode. link: http://identifiers.org/pubmed/11004387

Parameters:

NameDescription
Kch=4100.0; K1=5.0Reaction: CaER => Ca_cyt; Ca_cyt, Rate Law: Cytosol*Kch*Ca_cyt^2*(CaER-Ca_cyt)/(K1^2+Ca_cyt^2)
Kleak=0.05Reaction: CaER => Ca_cyt, Rate Law: Cytosol*Kleak*(CaER-Ca_cyt)
Kpump=20.0Reaction: Ca_cyt => CaER, Rate Law: Endoplasmic_Reticulum*Kpump*Ca_cyt
Kplus=0.1Reaction: Pr + Ca_cyt => CaPr, Rate Law: Cytosol*Kplus*Ca_cyt*Pr
Kin=300.0; K2=0.8Reaction: Ca_cyt => CaM; Ca_cyt, Rate Law: Mitochondria*Kin*Ca_cyt^8/(K2^8+Ca_cyt^8)
Kminus=0.01Reaction: CaPr => Pr + Ca_cyt, Rate Law: Cytosol*Kminus*CaPr
Kout=125.0; K3=5.0; Km=0.00625Reaction: CaM => Ca_cyt; Ca_cyt, Rate Law: Cytosol*CaM*(Kout*Ca_cyt^2/(K3^2+Ca_cyt^2)+Km)

States:

NameDescription
CaPr[calcium(2+); Protein]
Pr[Protein; protein polypeptide chain]
CaER[calcium(2+)]
Ca cyt[calcium(2+)]
CaM[calcium(2+)]

Markevich2004 - MAPK double phosphorylation, ordered Michaelis-Menton: BIOMD0000000027v0.0.1

Markevich2004 - MAPK double phosphorylation, ordered Michaelis-MentonThe model corresponds to the schemas 1 and 2 of Mar…

Details

Mitogen-activated protein kinase (MAPK) cascades can operate as bistable switches residing in either of two different stable states. MAPK cascades are often embedded in positive feedback loops, which are considered to be a prerequisite for bistable behavior. Here we demonstrate that in the absence of any imposed feedback regulation, bistability and hysteresis can arise solely from a distributive kinetic mechanism of the two-site MAPK phosphorylation and dephosphorylation. Importantly, the reported kinetic properties of the kinase (MEK) and phosphatase (MKP3) of extracellular signal-regulated kinase (ERK) fulfill the essential requirements for generating a bistable switch at a single MAPK cascade level. Likewise, a cycle where multisite phosphorylations are performed by different kinases, but dephosphorylation reactions are catalyzed by the same phosphatase, can also exhibit bistability and hysteresis. Hence, bistability induced by multisite covalent modification may be a widespread mechanism of the control of protein activity. link: http://identifiers.org/pubmed/14744999

Parameters:

NameDescription
Km2 = 500.0; k1cat = 0.01; Km1 = 50.0Reaction: M => Mp; MAPKK, Rate Law: uVol*k1cat*MAPKK*M/Km1/(1+M/Km1+Mp/Km2)
Km2 = 500.0; Km1 = 50.0; k2cat = 15.0Reaction: Mp => Mpp; MAPKK, M, Rate Law: uVol*k2cat*MAPKK*Mp/Km2/(1+M/Km1+Mp/Km2)
k4cat = 0.06; Km5 = 78.0; Km4 = 18.0; Km3 = 22.0Reaction: Mp => M; MKP3, Mpp, Rate Law: uVol*k4cat*MKP3*Mp/Km4/(1+Mpp/Km3+Mp/Km4+M/Km5)
Km5 = 78.0; Km4 = 18.0; Km3 = 22.0; k3cat = 0.084Reaction: Mpp => Mp; MKP3, M, Rate Law: uVol*k3cat*MKP3*Mpp/Km3/(1+Mpp/Km3+Mp/Km4+M/Km5)

States:

NameDescription
M[Mitogen-activated protein kinase 1]
Mp[Mitogen-activated protein kinase 1]
Mpp[Mitogen-activated protein kinase 1]

Markevich2004_MAPK_AllRandomElementary: BIOMD0000000030v0.0.1

This model originates from BioModels Database: A Database of Annotated Published Models. It is copyright (c) 2005-2009 T…

Details

Mitogen-activated protein kinase (MAPK) cascades can operate as bistable switches residing in either of two different stable states. MAPK cascades are often embedded in positive feedback loops, which are considered to be a prerequisite for bistable behavior. Here we demonstrate that in the absence of any imposed feedback regulation, bistability and hysteresis can arise solely from a distributive kinetic mechanism of the two-site MAPK phosphorylation and dephosphorylation. Importantly, the reported kinetic properties of the kinase (MEK) and phosphatase (MKP3) of extracellular signal-regulated kinase (ERK) fulfill the essential requirements for generating a bistable switch at a single MAPK cascade level. Likewise, a cycle where multisite phosphorylations are performed by different kinases, but dephosphorylation reactions are catalyzed by the same phosphatase, can also exhibit bistability and hysteresis. Hence, bistability induced by multisite covalent modification may be a widespread mechanism of the control of protein activity. link: http://identifiers.org/pubmed/14744999

Parameters:

NameDescription
k2 = 0.01Reaction: M_MAPKK_Y => MpY + MAPKK, Rate Law: cell*k2*M_MAPKK_Y
h7 = 0.01; h_7 = 1.0Reaction: MpY + MKP => MpY_MKP_Y, Rate Law: cell*(h7*MpY*MKP-h_7*MpY_MKP_Y)
h9 = 0.14; h_9 = 0.0018Reaction: M_MKP_Y => M + MKP, Rate Law: cell*(h9*M_MKP_Y-h_9*M*MKP)
k6 = 0.01Reaction: M_MAPKK_T => MpT + MAPKK, Rate Law: cell*k6*M_MAPKK_T
k_3 = 1.0; k3 = 0.032Reaction: MpY + MAPKK => MpY_MAPKK, Rate Law: cell*(k3*MpY*MAPKK-k_3*MpY_MAPKK)
k8 = 15.0Reaction: MpT_MAPKK => Mpp + MAPKK, Rate Law: cell*k8*MpT_MAPKK
h_1 = 1.0; h1 = 0.045Reaction: Mpp + MKP => Mpp_MKP_Y, Rate Law: cell*(h1*Mpp*MKP-h_1*Mpp_MKP_Y)
k4 = 15.0Reaction: MpY_MAPKK => Mpp + MAPKK, Rate Law: cell*k4*MpY_MAPKK
h2 = 0.092Reaction: Mpp_MKP_Y => MpT_MKP_Y, Rate Law: cell*h2*Mpp_MKP_Y
h3 = 1.0; h_3 = 0.01Reaction: MpT_MKP_Y => MpT + MKP, Rate Law: cell*(h3*MpT_MKP_Y-h_3*MpT*MKP)
h_12 = 0.01; h12 = 1.0Reaction: MpY_MKP_T => MpY + MKP, Rate Law: cell*(h12*MpY_MKP_T-h_12*MpY*MKP)
h_10 = 1.0; h10 = 0.045Reaction: Mpp + MKP => Mpp_MKP_T, Rate Law: cell*(h10*Mpp*MKP-h_10*Mpp_MKP_T)
k_7 = 1.0; k7 = 0.032Reaction: MpT + MAPKK => MpT_MAPKK, Rate Law: cell*(k7*MpT*MAPKK-k_7*MpT_MAPKK)
h11 = 0.092Reaction: Mpp_MKP_T => MpY_MKP_T, Rate Law: cell*h11*Mpp_MKP_T
h8 = 0.47Reaction: MpY_MKP_Y => M_MKP_Y, Rate Law: cell*h8*MpY_MKP_Y
k5 = 0.02; k_5 = 1.0Reaction: M + MAPKK => M_MAPKK_T, Rate Law: cell*(k5*M*MAPKK-k_5*M_MAPKK_T)
h6 = 0.086; h_6 = 0.0011Reaction: M_MKP_T => M + MKP, Rate Law: cell*(h6*M_MKP_T-h_6*M*MKP)
h4 = 0.01; h_4 = 1.0Reaction: MpT + MKP => MpT_MKP_T, Rate Law: cell*(h4*MpT*MKP-h_4*MpT_MKP_T)
k1 = 0.02; k_1 = 1.0Reaction: M + MAPKK => M_MAPKK_Y, Rate Law: cell*(k1*M*MAPKK-k_1*M_MAPKK_Y)
h5 = 0.5Reaction: MpT_MKP_T => M_MKP_T, Rate Law: cell*h5*MpT_MKP_T

States:

NameDescription
MAPKK[Dual specificity mitogen-activated protein kinase kinase 1]
MKP[Dual specificity protein phosphatase 1-B]
M MKP Y[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]
MpT[Mitogen-activated protein kinase 1]
MpT MKP T[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]
MpY MKP T[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]
Mpp MKP Y[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]
M MKP T[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]
M[Mitogen-activated protein kinase 1]
Mpp MKP T[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]
MpY MAPKK[Mitogen-activated protein kinase 1; Dual specificity mitogen-activated protein kinase kinase 1]
M MAPKK T[Dual specificity mitogen-activated protein kinase kinase 1; Mitogen-activated protein kinase 1]
MpT MKP Y[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]
MpY[Mitogen-activated protein kinase 1]
Mpp[Mitogen-activated protein kinase 1]
MpT MAPKK[Dual specificity mitogen-activated protein kinase kinase 1; Mitogen-activated protein kinase 1]
M MAPKK Y[Dual specificity mitogen-activated protein kinase kinase 1; Mitogen-activated protein kinase 1]
MpY MKP Y[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]

Markevich2004_MAPK_orderedElementary: BIOMD0000000026v0.0.1

The model corresponds to the schemas 1 and 2 of Markevich et al 2004, as described in the figure 1 and the supplementary…

Details

Mitogen-activated protein kinase (MAPK) cascades can operate as bistable switches residing in either of two different stable states. MAPK cascades are often embedded in positive feedback loops, which are considered to be a prerequisite for bistable behavior. Here we demonstrate that in the absence of any imposed feedback regulation, bistability and hysteresis can arise solely from a distributive kinetic mechanism of the two-site MAPK phosphorylation and dephosphorylation. Importantly, the reported kinetic properties of the kinase (MEK) and phosphatase (MKP3) of extracellular signal-regulated kinase (ERK) fulfill the essential requirements for generating a bistable switch at a single MAPK cascade level. Likewise, a cycle where multisite phosphorylations are performed by different kinases, but dephosphorylation reactions are catalyzed by the same phosphatase, can also exhibit bistability and hysteresis. Hence, bistability induced by multisite covalent modification may be a widespread mechanism of the control of protein activity. link: http://identifiers.org/pubmed/14744999

Parameters:

NameDescription
k1 = 0.02; k_1 = 1.0Reaction: M + MAPKK => M_MAPKK, Rate Law: uVol*(k1*M*MAPKK-k_1*M_MAPKK)
k2 = 0.01Reaction: M_MAPKK => Mp + MAPKK, Rate Law: uVol*k2*M_MAPKK
k_3 = 1.0; k3 = 0.032Reaction: Mp + MAPKK => Mp_MAPKK, Rate Law: uVol*(k3*Mp*MAPKK-k_3*Mp_MAPKK)
h_1 = 1.0; h1 = 0.045Reaction: Mpp + MKP3 => Mpp_MKP3, Rate Law: uVol*(h1*Mpp*MKP3-h_1*Mpp_MKP3)
k4 = 15.0Reaction: Mp_MAPKK => Mpp + MAPKK, Rate Law: uVol*k4*Mp_MAPKK
h2 = 0.092Reaction: Mpp_MKP3 => Mp_MKP3_dep, Rate Law: uVol*h2*Mpp_MKP3
h6 = 0.086; h_6 = 0.0011Reaction: M_MKP3 => M + MKP3, Rate Law: uVol*(h6*M_MKP3-h_6*M*MKP3)
h3 = 1.0; h_3 = 0.01Reaction: Mp_MKP3_dep => Mp + MKP3, Rate Law: h3*Mp_MKP3_dep-h_3*Mp*MKP3
h5 = 0.5Reaction: Mp_MKP3 => M_MKP3, Rate Law: uVol*h5*Mp_MKP3
h4 = 0.01; h_4 = 1.0Reaction: Mp + MKP3 => Mp_MKP3, Rate Law: uVol*(h4*Mp*MKP3-h_4*Mp_MKP3)

States:

NameDescription
MAPKK[Dual specificity mitogen-activated protein kinase kinase 1]
Mp MKP3 dep[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]
Mp MKP3[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]
Mpp MKP3[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]
Mp MAPKK[Dual specificity mitogen-activated protein kinase kinase 1; Mitogen-activated protein kinase 1]
M MAPKK[Mitogen-activated protein kinase 1; Dual specificity mitogen-activated protein kinase kinase 1]
M[Mitogen-activated protein kinase 1]
Mp[Mitogen-activated protein kinase 1]
M MKP3[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]
Mpp[Mitogen-activated protein kinase 1]
MKP3[Dual specificity protein phosphatase 1-B]

Markevich2004_MAPK_orderedMM2kinases: BIOMD0000000031v0.0.1

The model describes the double phosphorylation of MAP kinase by an ordered mechanism using the Michaelis-Menten formalis…

Details

Mitogen-activated protein kinase (MAPK) cascades can operate as bistable switches residing in either of two different stable states. MAPK cascades are often embedded in positive feedback loops, which are considered to be a prerequisite for bistable behavior. Here we demonstrate that in the absence of any imposed feedback regulation, bistability and hysteresis can arise solely from a distributive kinetic mechanism of the two-site MAPK phosphorylation and dephosphorylation. Importantly, the reported kinetic properties of the kinase (MEK) and phosphatase (MKP3) of extracellular signal-regulated kinase (ERK) fulfill the essential requirements for generating a bistable switch at a single MAPK cascade level. Likewise, a cycle where multisite phosphorylations are performed by different kinases, but dephosphorylation reactions are catalyzed by the same phosphatase, can also exhibit bistability and hysteresis. Hence, bistability induced by multisite covalent modification may be a widespread mechanism of the control of protein activity. link: http://identifiers.org/pubmed/14744999

Parameters:

NameDescription
Km2 = 500.0; k2cat = 15.0Reaction: Mp => Mpp; MAPKK2, M, Rate Law: uVol*k2cat*MAPKK2*Mp/Km2/(1+Mp/Km2)
k1cat = 0.01; Km1 = 50.0Reaction: M => Mp; MAPKK1, Rate Law: uVol*k1cat*MAPKK1*M/Km1/(1+M/Km1)
k4cat = 0.06; Km5 = 78.0; Km3 = 5.0; Km4 = 18.0Reaction: Mp => M; MKP3, Mpp, Rate Law: uVol*k4cat*MKP3*Mp/Km4/(1+Mpp/Km3+Mp/Km4+M/Km5)
Km5 = 78.0; Km3 = 5.0; Km4 = 18.0; k3cat = 0.084Reaction: Mpp => Mp; MKP3, M, Rate Law: uVol*k3cat*MKP3*Mpp/Km3/(1+Mpp/Km3+Mp/Km4+M/Km5)

States:

NameDescription
M[Mitogen-activated protein kinase 1]
Mp[Mitogen-activated protein kinase 1]
Mpp[Mitogen-activated protein kinase 1]

Markevich2004_MAPK_phosphoRandomElementary: BIOMD0000000028v0.0.1

The model corresponds to the schema 3 of Markevich et al 2004, as described in the figure 2 and the supplementary table…

Details

Mitogen-activated protein kinase (MAPK) cascades can operate as bistable switches residing in either of two different stable states. MAPK cascades are often embedded in positive feedback loops, which are considered to be a prerequisite for bistable behavior. Here we demonstrate that in the absence of any imposed feedback regulation, bistability and hysteresis can arise solely from a distributive kinetic mechanism of the two-site MAPK phosphorylation and dephosphorylation. Importantly, the reported kinetic properties of the kinase (MEK) and phosphatase (MKP3) of extracellular signal-regulated kinase (ERK) fulfill the essential requirements for generating a bistable switch at a single MAPK cascade level. Likewise, a cycle where multisite phosphorylations are performed by different kinases, but dephosphorylation reactions are catalyzed by the same phosphatase, can also exhibit bistability and hysteresis. Hence, bistability induced by multisite covalent modification may be a widespread mechanism of the control of protein activity. link: http://identifiers.org/pubmed/14744999

Parameters:

NameDescription
h6 = 0.086; h_6 = 0.0011Reaction: M_MKP3_T => M + MKP3, Rate Law: cell*(h6*M_MKP3_T-h_6*M*MKP3)
h9 = 0.14; h_9 = 0.0018Reaction: M_MKP3_Y => M + MKP3, Rate Law: cell*(h9*M_MKP3_Y-h_9*M*MKP3)
k3 = 0.025; k_3 = 1.0Reaction: MpY + MEK => MpY_MEK, Rate Law: cell*(k3*MpY*MEK-k_3*MpY_MEK)
h2 = 0.092Reaction: Mpp_MKP3 => MpT_MKP3_Y, Rate Law: cell*h2*Mpp_MKP3
h_1 = 1.0; h1 = 0.045Reaction: Mpp + MKP3 => Mpp_MKP3, Rate Law: cell*(h1*Mpp*MKP3-h_1*Mpp_MKP3)
h3 = 1.0; h_3 = 0.01Reaction: MpT_MKP3_Y => MpT + MKP3, Rate Law: cell*(h3*MpT_MKP3_Y-h_3*MpT*MKP3)
k_5 = 1.0; k5 = 0.05Reaction: M + MEK => M_MEK_T, Rate Law: cell*(k5*M*MEK-k_5*M_MEK_T)
h8 = 0.47Reaction: MpY_MKP3 => M_MKP3_Y, Rate Law: cell*h8*MpY_MKP3
k8 = 0.45Reaction: MpT_MEK => Mpp + MEK, Rate Law: cell*k8*MpT_MEK
k1 = 0.005; k_1 = 1.0Reaction: M + MEK => M_MEK_Y, Rate Law: cell*(k1*M*MEK-k_1*M_MEK_Y)
k_7 = 1.0; k7 = 0.005Reaction: MpT + MEK => MpT_MEK, Rate Law: cell*(k7*MpT*MEK-k_7*MpT_MEK)
k6 = 0.008Reaction: M_MEK_T => MpT + MEK, Rate Law: cell*k6*M_MEK_T
k4 = 0.007Reaction: MpY_MEK => Mpp + MEK, Rate Law: cell*k4*MpY_MEK
h7 = 0.01; h_7 = 1.0Reaction: MpY + MKP3 => MpY_MKP3, Rate Law: cell*(h7*MpY*MKP3-h_7*MpY_MKP3)
h4 = 0.01; h_4 = 1.0Reaction: MpT + MKP3 => MpT_MKP3_T, Rate Law: cell*(h4*MpT*MKP3-h_4*MpT_MKP3_T)
h5 = 0.5Reaction: MpT_MKP3_T => M_MKP3_T, Rate Law: cell*h5*MpT_MKP3_T
k2 = 1.08Reaction: M_MEK_Y => MpY + MEK, Rate Law: cell*k2*M_MEK_Y

States:

NameDescription
MpT MKP3 Y[Mitogen-activated protein kinase 1; Dual specificity protein phosphatase 1-B]
M MEK Y[Mitogen-activated protein kinase 1; Dual specificity mitogen-activated protein kinase kinase 1]
Mpp MKP3[Mitogen-activated protein kinase 1; Dual specificity protein phosphatase 1-B]
MEK[Dual specificity mitogen-activated protein kinase kinase 1]
MpY MKP3[Mitogen-activated protein kinase 1; Dual specificity protein phosphatase 1-B]
M MEK T[Mitogen-activated protein kinase 1; Dual specificity mitogen-activated protein kinase kinase 1]
MpY MEK[Dual specificity mitogen-activated protein kinase kinase 1; Mitogen-activated protein kinase 1]
MpT[Mitogen-activated protein kinase 1]
MpT MKP3 T[Mitogen-activated protein kinase 1; Dual specificity protein phosphatase 1-B]
M MKP3 Y[Dual specificity protein phosphatase 1-B; Mitogen-activated protein kinase 1]
M MKP3 T[Mitogen-activated protein kinase 1; Dual specificity protein phosphatase 1-B]
M[Mitogen-activated protein kinase 1]
MpT MEK[Mitogen-activated protein kinase 1; Dual specificity mitogen-activated protein kinase kinase 1]
MpY[Mitogen-activated protein kinase 1]
Mpp[Mitogen-activated protein kinase 1]
MKP3[Dual specificity protein phosphatase 1-B]

Markevich2004_MAPK_phosphoRandomMM: BIOMD0000000029v0.0.1

The model corresponds to the schema 3 of Markevich et al 2004, as described in the figure 2 and the supplementary table…

Details

Mitogen-activated protein kinase (MAPK) cascades can operate as bistable switches residing in either of two different stable states. MAPK cascades are often embedded in positive feedback loops, which are considered to be a prerequisite for bistable behavior. Here we demonstrate that in the absence of any imposed feedback regulation, bistability and hysteresis can arise solely from a distributive kinetic mechanism of the two-site MAPK phosphorylation and dephosphorylation. Importantly, the reported kinetic properties of the kinase (MEK) and phosphatase (MKP3) of extracellular signal-regulated kinase (ERK) fulfill the essential requirements for generating a bistable switch at a single MAPK cascade level. Likewise, a cycle where multisite phosphorylations are performed by different kinases, but dephosphorylation reactions are catalyzed by the same phosphatase, can also exhibit bistability and hysteresis. Hence, bistability induced by multisite covalent modification may be a widespread mechanism of the control of protein activity. link: http://identifiers.org/pubmed/14744999

Parameters:

NameDescription
Km7 = 34.0; kcat5 = 0.084; Km6 = 18.0; Km8 = 40.0; Km5 = 22.0Reaction: Mpp => MpT; MKP3, MpY, M, Rate Law: cell*kcat5*MKP3*Mpp/Km5/(1+Mpp/Km5+MpT/Km6+MpY/Km7+M/Km8)
kcat4 = 0.45; Km4 = 300.0; Km2 = 40.0; Km1 = 410.0; Km3 = 20.0Reaction: MpT => Mpp; MEK, M, MpY, Rate Law: cell*kcat4*MEK*MpT/Km4/(1+M*(Km1+Km3)/(Km1*Km3)+MpY/Km2+MpT/Km4)
Km7 = 34.0; kcat7 = 0.108; Km6 = 18.0; Km8 = 40.0; Km5 = 22.0Reaction: MpY => M; MKP3, Mpp, MpT, Rate Law: cell*kcat7*MKP3*MpY/Km7/(1+Mpp/Km5+MpT/Km6+MpY/Km7+M/Km8)
kcat3 = 0.008; Km4 = 300.0; Km2 = 40.0; Km3 = 20.0; Km1 = 410.0Reaction: M => MpT; MEK, MpY, Rate Law: cell*kcat3*MEK*M/Km3/(1+M*(Km1+Km3)/(Km1*Km3)+MpY/Km2+MpT/Km4)
Km7 = 34.0; kcat6 = 0.06; Km6 = 18.0; Km8 = 40.0; Km5 = 22.0Reaction: MpT => M; MKP3, Mpp, MpY, Rate Law: cell*kcat6*MKP3*MpT/Km6/(1+Mpp/Km5+MpT/Km6+MpY/Km7+M/Km8)
Km4 = 300.0; kcat1 = 1.08; Km2 = 40.0; Km1 = 410.0; Km3 = 20.0Reaction: M => MpY; MEK, MpT, Rate Law: cell*kcat1*MEK*M/Km1/(1+M*(Km1+Km3)/(Km1*Km3)+MpY/Km2+MpT/Km4)
kcat2 = 0.007; Km4 = 300.0; Km2 = 40.0; Km1 = 410.0; Km3 = 20.0Reaction: MpY => Mpp; MEK, M, MpT, Rate Law: cell*kcat2*MEK*MpY/Km2/(1+M*(Km1+Km3)/(Km1*Km3)+MpY/Km2+MpT/Km4)

States:

NameDescription
M[Mitogen-activated protein kinase 1]
MpY[Mitogen-activated protein kinase 1]
Mpp[Mitogen-activated protein kinase 1]
MpT[Mitogen-activated protein kinase 1]

Martinez-Guimera2017 - Generic negative feedback circuit (Model 4): MODEL1710260003v0.0.1

Martinez-Guimera2017 - Generic negative feedback circuit (Model 4)This model is described in the article: ['Molecular h…

Details

The ability of reactive oxygen species (ROS) to cause molecular damage has meant that chronic oxidative stress has been mostly studied from the point of view of being a source of toxicity to the cell. However, the known duality of ROS molecules as both damaging agents and cellular redox signals implies another perspective in the study of sustained oxidative stress. This is a perspective of studying oxidative stress as a constitutive signal within the cell. In this work, we adopt a theoretical perspective as an exploratory and explanatory approach to examine how chronic oxidative stress can interfere with signal processing by redox signalling pathways in the cell. We report that constitutive signals can give rise to a 'molecular habituation' effect that can prime for a gradual loss of biological function. This is because a constitutive signal in the environment has the potential to reduce the responsiveness of a signalling pathway through the prolonged activation of negative regulators. Additionally, we demonstrate how this phenomenon is likely to occur in different signalling pathways exposed to persistent signals and furthermore at different levels of biological organisation. link: http://identifiers.org/pubmed/29146308

Martinez-Guimera2017 - Generic negative feedforward circuit (Model 5): MODEL1710260004v0.0.1

Martinez-Guimera2017 - Generic negative feedforward circuit (Model 5)This model is described in the article: ['Molecula…

Details

The ability of reactive oxygen species (ROS) to cause molecular damage has meant that chronic oxidative stress has been mostly studied from the point of view of being a source of toxicity to the cell. However, the known duality of ROS molecules as both damaging agents and cellular redox signals implies another perspective in the study of sustained oxidative stress. This is a perspective of studying oxidative stress as a constitutive signal within the cell. In this work, we adopt a theoretical perspective as an exploratory and explanatory approach to examine how chronic oxidative stress can interfere with signal processing by redox signalling pathways in the cell. We report that constitutive signals can give rise to a 'molecular habituation' effect that can prime for a gradual loss of biological function. This is because a constitutive signal in the environment has the potential to reduce the responsiveness of a signalling pathway through the prolonged activation of negative regulators. Additionally, we demonstrate how this phenomenon is likely to occur in different signalling pathways exposed to persistent signals and furthermore at different levels of biological organisation. link: http://identifiers.org/pubmed/29146308

Martinez-Guimera2017 - Generic redox signalling model with negative feedback regulation (Model 2): MODEL1710260001v0.0.1

Martinez-Guimera2017 - Generic redox signalling model with negative feedback regulation (Model 2)This model is described…

Details

The ability of reactive oxygen species (ROS) to cause molecular damage has meant that chronic oxidative stress has been mostly studied from the point of view of being a source of toxicity to the cell. However, the known duality of ROS molecules as both damaging agents and cellular redox signals implies another perspective in the study of sustained oxidative stress. This is a perspective of studying oxidative stress as a constitutive signal within the cell. In this work, we adopt a theoretical perspective as an exploratory and explanatory approach to examine how chronic oxidative stress can interfere with signal processing by redox signalling pathways in the cell. We report that constitutive signals can give rise to a 'molecular habituation' effect that can prime for a gradual loss of biological function. This is because a constitutive signal in the environment has the potential to reduce the responsiveness of a signalling pathway through the prolonged activation of negative regulators. Additionally, we demonstrate how this phenomenon is likely to occur in different signalling pathways exposed to persistent signals and furthermore at different levels of biological organisation. link: http://identifiers.org/pubmed/29146308

Martinez-Guimera2017 - Generic redox signalling model with negative feedback regulation (Model 2): MODEL2001080001v0.0.1

Martinez-Guimera2017 - Generic redox signalling model with negative feedback regulation (Model 2)This model is described…

Details

The ability of reactive oxygen species (ROS) to cause molecular damage has meant that chronic oxidative stress has been mostly studied from the point of view of being a source of toxicity to the cell. However, the known duality of ROS molecules as both damaging agents and cellular redox signals implies another perspective in the study of sustained oxidative stress. This is a perspective of studying oxidative stress as a constitutive signal within the cell. In this work, we adopt a theoretical perspective as an exploratory and explanatory approach to examine how chronic oxidative stress can interfere with signal processing by redox signalling pathways in the cell. We report that constitutive signals can give rise to a 'molecular habituation' effect that can prime for a gradual loss of biological function. This is because a constitutive signal in the environment has the potential to reduce the responsiveness of a signalling pathway through the prolonged activation of negative regulators. Additionally, we demonstrate how this phenomenon is likely to occur in different signalling pathways exposed to persistent signals and furthermore at different levels of biological organisation. link: http://identifiers.org/pubmed/29146308

Martinez-Guimera2017 - Generic redox signalling model with negative feedforward regulation (Model 3): MODEL1710260002v0.0.1

Martinez-Guimera2017 - Generic redox signalling model with negative feedforward regulation (Model 3)This model is descri…

Details

The ability of reactive oxygen species (ROS) to cause molecular damage has meant that chronic oxidative stress has been mostly studied from the point of view of being a source of toxicity to the cell. However, the known duality of ROS molecules as both damaging agents and cellular redox signals implies another perspective in the study of sustained oxidative stress. This is a perspective of studying oxidative stress as a constitutive signal within the cell. In this work, we adopt a theoretical perspective as an exploratory and explanatory approach to examine how chronic oxidative stress can interfere with signal processing by redox signalling pathways in the cell. We report that constitutive signals can give rise to a 'molecular habituation' effect that can prime for a gradual loss of biological function. This is because a constitutive signal in the environment has the potential to reduce the responsiveness of a signalling pathway through the prolonged activation of negative regulators. Additionally, we demonstrate how this phenomenon is likely to occur in different signalling pathways exposed to persistent signals and furthermore at different levels of biological organisation. link: http://identifiers.org/pubmed/29146308

Martinez-Guimera2017 - Generic redox signalling model without negative regulation (Model 1): MODEL1710260000v0.0.1

Martinez-Guimera2017 - Generic redox signalling model without negative regulation (Model 1)This model is described in th…

Details

The ability of reactive oxygen species (ROS) to cause molecular damage has meant that chronic oxidative stress has been mostly studied from the point of view of being a source of toxicity to the cell. However, the known duality of ROS molecules as both damaging agents and cellular redox signals implies another perspective in the study of sustained oxidative stress. This is a perspective of studying oxidative stress as a constitutive signal within the cell. In this work, we adopt a theoretical perspective as an exploratory and explanatory approach to examine how chronic oxidative stress can interfere with signal processing by redox signalling pathways in the cell. We report that constitutive signals can give rise to a 'molecular habituation' effect that can prime for a gradual loss of biological function. This is because a constitutive signal in the environment has the potential to reduce the responsiveness of a signalling pathway through the prolonged activation of negative regulators. Additionally, we demonstrate how this phenomenon is likely to occur in different signalling pathways exposed to persistent signals and furthermore at different levels of biological organisation. link: http://identifiers.org/pubmed/29146308

Martinez-Sanchez2015 - T CD4+ lymphocyte transcriptional-signaling regulatory network: BIOMD0000000593v0.0.1

Martinez-Sanchez2015 - T CD4+ lymphocyte transcriptional-signaling regulatory networkThis model is described in the arti…

Details

CD4+ T cells orchestrate the adaptive immune response in vertebrates. While both experimental and modeling work has been conducted to understand the molecular genetic mechanisms involved in CD4+ T cell responses and fate attainment, the dynamic role of intrinsic (produced by CD4+ T lymphocytes) versus extrinsic (produced by other cells) components remains unclear, and the mechanistic and dynamic understanding of the plastic responses of these cells remains incomplete. In this work, we studied a regulatory network for the core transcription factors involved in CD4+ T cell-fate attainment. We first show that this core is not sufficient to recover common CD4+ T phenotypes. We thus postulate a minimal Boolean regulatory network model derived from a larger and more comprehensive network that is based on experimental data. The minimal network integrates transcriptional regulation, signaling pathways and the micro-environment. This network model recovers reported configurations of most of the characterized cell types (Th0, Th1, Th2, Th17, Tfh, Th9, iTreg, and Foxp3-independent T regulatory cells). This transcriptional-signaling regulatory network is robust and recovers mutant configurations that have been reported experimentally. Additionally, this model recovers many of the plasticity patterns documented for different T CD4+ cell types, as summarized in a cell-fate map. We tested the effects of various micro-environments and transient perturbations on such transitions among CD4+ T cell types. Interestingly, most cell-fate transitions were induced by transient activations, with the opposite behavior associated with transient inhibitions. Finally, we used a novel methodology was used to establish that T-bet, TGF-β and suppressors of cytokine signaling proteins are keys to recovering observed CD4+ T cell plastic responses. In conclusion, the observed CD4+ T cell-types and transition patterns emerge from the feedback between the intrinsic or intracellular regulatory core and the micro-environment. We discuss the broader use of this approach for other plastic systems and possible therapeutic interventions. link: http://identifiers.org/pubmed/26090929

Martins2001_glyoxalase: MODEL6624199343v0.0.1

. . . **[SBML](http://www.sbml.org/) level 2 code generated for the JWS Online project by Jacky Snoep using [PySCeS]…

Details

The kinetics of glyoxalase I [(R)-S-lactoylglutathione methylglyoxal-lyase; EC 4.4.1.5] and glyoxalase II (S-2-hydroxyacylglutathione hydrolase; EC 3.1.2.6) from Saccharomyces cerevisiae was studied in situ, in digitonin permeabilized cells, using two different approaches: initial rate analysis and progress curves analysis. Initial rate analysis was performed by hyperbolic regression of initial rates using the program HYPERFIT. Glyoxalase I exhibited saturation kinetics on 0.05-2.5 mM hemithioacetal concentration range, with kinetic parameters Km 0.53 +/- 0.07 mM and V (3.18 +/- 0.16) x 10(-2) mM.min(-1). Glyoxalase II also showed saturation kinetics in the SD-lactoylglutathione concentration range of 0.15-3 mM and Km 0.32 +/- 0.13 mM and V (1.03 +/- 0.10) x 10(-3) mM.min(-1) were obtained. The kinetic parameters of both enzymes were also estimated by nonlinear regression of progress curves using the raw absorbance data and integrated differential rate equations with the program GEPASI. Several optimization methods were used to minimize the sum of squares of residuals. The best parameter fit for the glyoxalase I reaction was obtained with a single curve analysis, using the irreversible Michaelis-Menten model. The kinetic parameters obtained, Km 0.62 +/- 0.18 mM and V (2.86 +/- 0.01) x 10(-2) mM.min(-1), were in agreement with those obtained by initial rate analysis. The results obtained for glyoxalase II, using either the irreversible Michaelis-Menten model or a phenomenological reversible hyperbolic model, showed a high correlation of residuals with time and/or high values of standard deviation associated with Km. The possible causes for the discrepancy between data obtained from initial rate analysis and progress curve analysis, for glyoxalase II, are discussed. link: http://identifiers.org/pubmed/11453985

Martins2003_AmadoriDegradation: BIOMD0000000050v0.0.1

This a model from the article: Kinetic modelling of Amadori N-(1-deoxy-D-fructos-1-yl)-glycine degradation pathways.…

Details

A kinetic model for N-(1-deoxy-D-fructos-1-yl)-glycine (DFG) thermal decomposition was proposed. Two temperatures (100 and 120 degrees C) and two pHs (5.5 and 6.8) were studied. The measured responses were DFG, 3-deoxyosone, 1-deoxyosone, methylglyoxal, acetic acid, formic acid, glucose, fructose, mannose and melanoidins. For each system the model parameters, the rate constants, were estimated by non-linear regression, via multiresponse modelling. The determinant criterion was used as the statistical fit criterion. Model discrimination was performed by both chemical insight and statistical tests (Posterior Probability and Akaike criterion). Kinetic analysis showed that at lower pH DFG 1,2-enolization is favoured whereas with increasing pH 2,3-enolization becomes a more relevant degradation pathway. The lower amount observed of 1-DG is related with its high reactivity. It was shown that acetic acid, a main degradation product from DFG, was mainly formed through 1-DG degradation. Also from the estimated parameters 3-DG was found to be the main precursor in carbohydrate fragments formation, responsible for colour formation. Some indication was given that as the reaction proceeded other compounds besides DFG become reactants themselves with the formation among others of methylglyoxal. The multiresponse kinetic analysis was shown to be both helpful in deriving relevant kinetic parameters as well as in obtaining insight into the reaction mechanism. link: http://identifiers.org/pubmed/12873422

Parameters:

NameDescription
k3=0.0155Reaction: DFG => Gly + Cn, Rate Law: k3*DFG
k6=0.0274Reaction: _3DG => FA, Rate Law: k6*_3DG
k13=0.0022Reaction: Glu => _3DG, Rate Law: k13*Glu
k1=0.0057Reaction: DFG => E1, Rate Law: k1*DFG
k10=0.0707Reaction: E1 => Gly + Man, Rate Law: k10*E1
k9=1.9085Reaction: _1DG => AA, Rate Law: k9*_1DG
k15=0.0159Reaction: Cn => AA + FA + MG, Rate Law: k15*Cn
k16=0.0134Reaction: E2 => Gly + Fru, Rate Law: k16*E2
k14=0.0034Reaction: Gly + Cn => Mel, Rate Law: k14*Cn*Gly
k4=0.0794Reaction: E1 => Gly + _3DG, Rate Law: k4*E1
k5=0.0907Reaction: _3DG => Cn, Rate Law: k5*_3DG
k12=8.0E-4Reaction: Man => Glu, Rate Law: k12*Man
k11=0.1131Reaction: E1 => Gly + Glu, Rate Law: k11*E1
k2=0.0156Reaction: DFG => E2, Rate Law: k2*DFG
k7=0.2125Reaction: E2 => Gly + _1DG, Rate Law: k7*E2
k8=0.0Reaction: _1DG => Cn, Rate Law: k8*_1DG

States:

NameDescription
Gly[glycine; Glycine]
MG[methylglyoxal; Methylglyoxal]
E2E2
Man[CHEBI_14575; D-Mannose]
FA[formic acid; Formate]
DFGDFG
MelMel
1DG_1DG
Cn[CHEBI_23008]
Glu[glucose; C00293]
AA[acetic acid; Acetate]
E1E1
Fru[fructose; Fructose]
3DG_3DG

Martins2004_Yeast_Glycolysis_GlycerolSynthesis: MODEL1009220000v0.0.1

**A systems biology study of two distinct growth phases of *Saccharomyces cerevisiae* cultures ** AM Martins, D Camac…

Details

Saccharomyces cerevisiae cultures growing exponentially and after starvation are distinctly different, as shown by several studies at the physiological, biochemical, and morphological levels. One group of studies attempted to be mechanistic, characterizing a few molecules and interactions, while another focused on global observations but remained descriptive or at best phenomenological. Recent advances in large-scale molecular profiling technologies, theoretical, and computational biology, are making possible integrative studies of biological systems, where global observations are explained through computational models with solid theoretical bases. A case study of the systems biology approach applied to the characterization of baker's yeast cultures in exponential growth and post-diauxic phases is presented.

Twenty cell cultures of S. cerevisiae were grown under similar environmental conditions. Samples from ten of these cultures were collected 11 hours after inoculation, while samples from the other ten were collected 4 days after inoculation. These samples were analyzed for their RNA and metabolite composition using Affymetrix chips and gas chromatography-mass spectrometry (GC-MS). The data were interpreted with statistical analyses and through the use of computer simulations of a kinetic model that was built by merging two independent models of glycolysis and glycerol biosynthesis. The simulation results agree with the exponential growth phase data, while no model is available for the post-diauxic phase. We discuss the need for expanding the number of kinetic models of S. cerevisiae, combining metabolism and genetic regulation, and covering all of its biochemistry. link: http://identifiers.org/doi/10.2174/1389202043348643

Martins2013 - True and apparent inhibition of amyloid fribril formation: BIOMD0000000561v0.0.1

Martins2013 - True and apparent inhibition of amyloid fribril formationThis model is described in the article: [True an…

Details

A possible therapeutic strategy for amyloid diseases involves the use of small molecule compounds to inhibit protein assembly into insoluble aggregates. According to the recently proposed Crystallization-Like Model, the kinetics of amyloid fibrillization can be retarded by decreasing the frequency of new fibril formation or by decreasing the elongation rate of existing fibrils. To the compounds that affect the nucleation and/or the growth steps we call true inhibitors. An apparent inhibition mechanism may however result from the alteration of thermodynamic properties such as the solubility of the amyloidogenic protein. Apparent inhibitors markedly influence protein aggregation kinetics measured in vitro, yet they are likely to lead to disappointing results when tested in vivo. This is because cells and tissues media are in general much more buffered against small variations in composition than the solutions prepared in lab. Here we show how to discriminate between true and apparent inhibition mechanisms from experimental data on protein aggregation kinetics. The goal is to be able to identify false positives much earlier during the drug development process. link: http://identifiers.org/pubmed/23232498

Parameters:

NameDescription
deltamt = 1.0; ka = 0.5; kb = 0.001Reaction: Amyloid = (1-1/(kb*(exp(ka*time)-1)+1))*deltamt, Rate Law: missing

States:

NameDescription
Amyloid[amyloid fibril]

MartínJiménez2017 - Genome-scale reconstruction of the human astrocyte metabolic network: MODEL1608180000v0.0.1

MartínJiménez2017 - Genome-scale reconstruction of the human astrocyte metabolic networkThis model is described in the a…

Details

Astrocytes are the most abundant cells of the central nervous system; they have a predominant role in maintaining brain metabolism. In this sense, abnormal metabolic states have been found in different neuropathological diseases. Determination of metabolic states of astrocytes is difficult to model using current experimental approaches given the high number of reactions and metabolites present. Thus, genome-scale metabolic networks derived from transcriptomic data can be used as a framework to elucidate how astrocytes modulate human brain metabolic states during normal conditions and in neurodegenerative diseases. We performed a Genome-Scale Reconstruction of the Human Astrocyte Metabolic Network with the purpose of elucidating a significant portion of the metabolic map of the astrocyte. This is the first global high-quality, manually curated metabolic reconstruction network of a human astrocyte. It includes 5,007 metabolites and 5,659 reactions distributed among 8 cell compartments, (extracellular, cytoplasm, mitochondria, endoplasmic reticle, Golgi apparatus, lysosome, peroxisome and nucleus). Using the reconstructed network, the metabolic capabilities of human astrocytes were calculated and compared both in normal and ischemic conditions. We identified reactions activated in these two states, which can be useful for understanding the astrocytic pathways that are affected during brain disease. Additionally, we also showed that the obtained flux distributions in the model, are in accordance with literature-based findings. Up to date, this is the most complete representation of the human astrocyte in terms of inclusion of genes, proteins, reactions and metabolic pathways, being a useful guide for in-silico analysis of several metabolic behaviors of the astrocyte during normal and pathologic states. link: http://identifiers.org/pubmed/28243200

Marwan2003 - Genetics, regulatory hierarchy between genes: BIOMD0000000037v0.0.1

To the extent possible under law, all copyright and related or neighbouring rights to this encoded model have been dedic…

Details

Mutants of Physarum polycephalum can be complemented by fusion of plasmodial cells followed by cytoplasmic mixing. Complementation between strains carrying different mutational defects in the sporulation control network may depend on the signaling state of the network components. We have previously suggested that time-resolved somatic complementation (TRSC) analysis with such mutants may be used to probe network architecture and dynamics. By computer simulation it is now shown how and under which conditions the regulatory hierarchy of genes can be determined experimentally. A kinetic model of the sporulation control network is developed, which is then used to demonstrate how the mechanisms of TRSC can be understood and simulated at the kinetic level. On the basis of theoretical considerations, experimental parameters that determine whether functional complementation of two mutations will occur are identified. It is also shown how gene dosage-effect relationships can be employed for network analysis. The theoretical framework provided may be used to systematically analyze network structure and dynamics through time-resolved somatic complementation studies. The conclusions drawn are of general relevance in that they do not depend on the validity of the model from which they were derived. link: http://identifiers.org/pubmed/12750324

Parameters:

NameDescription
kG=0.1Reaction: Ya + Gluc => Yi, Rate Law: kG*Ya*Gluc*compartment
alpha2=50.0Reaction: => V; S, Rate Law: compartment*alpha2/(1+S^3)
kd=0.1Reaction: Pr => Pi, Rate Law: compartment*kd*Pr
IfrSfrPfr=0.1Reaction: Pfr => Pr, Rate Law: compartment*Pfr*IfrSfrPfr
IrSrPr=0.0Reaction: Pr => Pfr, Rate Law: IrSrPr*Pr*compartment
kai=0.8Reaction: Xa => Xi, Rate Law: kai*Xa*compartment
ky=1.0Reaction: preS => S; Ya, Rate Law: preS*ky*Ya*compartment
alpha1=30.0Reaction: => S; V, Rate Law: compartment*alpha1/(1+V^3)
kd_v=1.0Reaction: V =>, Rate Law: compartment*V*kd_v
kia=0.1Reaction: Xi => Xa; Pr, Rate Law: Xi*kia*Pr*compartment
kd_s=1.0Reaction: S =>, Rate Law: kd_s*S*compartment
kx=0.2Reaction: prepreS => preS; Xa, Rate Law: prepreS*kx*Xa*compartment

States:

NameDescription
prepreSprepreS
XiXi
YaYa
YiYi
VV
Pfr[bilin; IPR001294]
XaXa
preSpreS
Pr[bilin; IPR001294]
SS
PiPi
Gluc[glucose; C00293]

Masel2000 - Drugs to stop prion aggregates and other amyloids: MODEL1410310000v0.0.1

Masel2000 - Drugs to stop prion aggregates and other amyloidsEncoded non-curated model. Issues:  - Missing initial conc…

Details

Amyloid protein aggregates are implicated in many neurodegenerative diseases, including Alzheimer's disease and the prion diseases. Therapeutics to block amyloid formation are often tested in vitro, but it is not clear how to extrapolate from these experiments to a clinical setting, where the effective drug dose may be much lower. Here we address this question using a theoretical kinetic model to calculate the growth rate of protein aggregates as a function of the dose of each of three categories of drug. We find that therapeutics which block the growing ends of amyloids are the most promising, as alternative strategies may be ineffective or even accelerate amyloid formation at low drug concentrations. Our mathematical model can be used to identify and optimise an end-blocking drug in vitro. Our model also suggests an alternative explanation for data previously thought to prove the existence of an entity known as protein X. link: http://identifiers.org/pubmed/11152275

Matsuoka2003_VentricularCells_SinoatrialNodePacemaker: MODEL1006230058v0.0.1

This a model from the article: Role of individual ionic current systems in ventricular cells hypothesized by a model s…

Details

Individual ion channels or exchangers are described with a common set of equations for both the sinoatrial node pacemaker and ventricular cells. New experimental data are included, such as the new kinetics of the inward rectifier K+ channel, delayed rectifier K+ channel, and sustained inward current. The gating model of Shirokov et al. (J Gen Physiol 102: 1005-1030, 1993) is used for both the fast Na+ and L-type Ca2+ channels. When combined with a contraction model (Negroni and Lascano: J Mol Cell Cardiol 28: 915-929, 1996), the experimental staircase phenomenon of contraction is reconstructed. The modulation of the action potential by varying the external Ca2+ and K+ concentrations is well simulated. The conductance of I(CaL) dominates membrane conductance during the action potential so that an artificial increase of I(to), I(Kr), I(Ks), or I(KATP) magnifies I(CaL) amplitude. Repolarizing current is provided sequentially by I(Ks), I(Kr), and I(K1). Depression of ATP production results in the shortening of action potential through the activation of I(KATP). The ratio of Ca2+ released from SR over Ca2+ entering via I(CaL) (Ca2+ gain = approximately 15) in excitation-contraction coupling well agrees with the experimental data. The model serves as a predictive tool in generating testable hypotheses. link: http://identifiers.org/pubmed/12877767

Maurya2005_GTPaseCycle_reducedOrder: BIOMD0000000085v0.0.1

This model is according to the paper Reduced-order modeling of biochemical networks: application to the GTPase-cycle sig…

Details

Biochemical systems embed complex networks and hence development and analysis of their detailed models pose a challenge for computation. Coarse-grained biochemical models, called reduced-order models (ROMs), consisting of essential biochemical mechanisms are more useful for computational analysis and for studying important features of a biochemical network. The authors present a novel method to model-reduction by identifying potentially important parameters using multidimensional sensitivity analysis. A ROM is generated for the GTPase-cycle module of m1 muscarinic acetylcholine receptor, Gq, and regulator of G-protein signalling 4 (a GTPase-activating protein or GAP) starting from a detailed model of 48 reactions. The resulting ROM has only 17 reactions. The ROM suggested that complexes of G-protein coupled receptor (GPCR) and GAP–which were proposed in the detailed model as a hypothesis–are required to fit the experimental data. Models previously published in the literature are also simulated and compared with the ROM. Through this comparison, a minimal ROM, that also requires complexes of GPCR and GAP, with just 15 parameters is generated. The proposed reduced-order modelling methodology is scalable to larger networks and provides a general framework for the reduction of models of biochemical systems. link: http://identifiers.org/pubmed/16986265

Parameters:

NameDescription
k2=0.00297; k1=25.0Reaction: species_15 => species_16 + species_7, Rate Law: compartment_0*(k1*species_15-k2*species_16*species_7)
k2=1.28; k1=1.32E8Reaction: species_5 + species_4 => species_10, Rate Law: compartment_0*(k1*species_5*species_4-k2*species_10)
k1=3.96E9; k2=5.43E-5Reaction: species_2 + species_4 => species_14, Rate Law: compartment_0*(k1*species_2*species_4-k2*species_14)
k1=853000.0; k2=0.00468Reaction: species_9 + species_3 => species_10, Rate Law: compartment_0*(k1*species_9*species_3-k2*species_10)
k2=2.22E-9; k1=0.013Reaction: species_10 => species_13 + species_7, Rate Law: compartment_0*(k1*species_10-k2*species_13*species_7)
k2=9.03E-7; k1=0.013Reaction: species_5 => species_6 + species_7, Rate Law: compartment_0*(k1*species_5-k2*species_6*species_7)
k1=386000.0; k2=0.0408Reaction: species_5 + species_0 => species_11, Rate Law: compartment_0*(k1*species_5*species_0-k2*species_11)
k2=0.478; k1=6300000.0Reaction: species_10 + species_0 => species_15, Rate Law: compartment_0*(k1*species_10*species_0-k2*species_15)
k1=2.0; k2=1470000.0Reaction: species_13 => species_9 + species_8, Rate Law: compartment_0*(k1*species_13-k2*species_9*species_8)
k2=2940.0; k1=2.75Reaction: species_16 => species_14 + species_8, Rate Law: compartment_0*(k1*species_16-k2*species_14*species_8)
k2=8.38E-6; k1=529000.0Reaction: species_1 + species_3 => species_5, Rate Law: compartment_0*(k1*species_1*species_3-k2*species_5)
k1=1.0E-4; k2=3.83Reaction: species_12 => species_2 + species_8, Rate Law: compartment_0*(k1*species_12-k2*species_2*species_8)
k1=64100.0; k2=0.95Reaction: species_6 + species_0 => species_12, Rate Law: compartment_0*(k1*species_6*species_0-k2*species_12)
k1=1620000.0; k2=0.00875Reaction: species_14 + species_3 => species_15, Rate Law: compartment_0*(k1*species_14*species_3-k2*species_15)
k1=9.47E7; k2=0.00227Reaction: species_6 + species_4 => species_13, Rate Law: compartment_0*(k1*species_6*species_4-k2*species_13)
k2=62.3; k1=1.0E-4Reaction: species_6 => species_1 + species_8, Rate Law: compartment_0*(k1*species_6-k2*species_1*species_8)
k1=25.0; k2=0.244Reaction: species_11 => species_12 + species_7, Rate Law: compartment_0*(k1*species_11-k2*species_12*species_7)

States:

NameDescription
species 9[heterotrimeric G-protein complex; receptor complex]
species 2[IPR000342; heterotrimeric G-protein complex]
species 6[GDP; heterotrimeric G-protein complex]
species 10[GTP; heterotrimeric G-protein complex; receptor complex]
species 11[GTP; IPR000342; heterotrimeric G-protein complex]
species 1[heterotrimeric G-protein complex]
species 4[receptor complex; IPR000337]
species 16[GDP; IPR000342; heterotrimeric G-protein complex; receptor complex]
species 14[IPR000342; heterotrimeric G-protein complex; receptor complex]
species 3[GTP; GTP]
species 0[IPR000342]
species 8[GDP; GDP]
species 12[GDP; IPR000342; heterotrimeric G-protein complex]
species 7[phosphate(3-)]
species 5[GTP; heterotrimeric G-protein complex]
species 15[GTP; IPR000342; heterotrimeric G-protein complex; receptor complex]
species 13[GDP; heterotrimeric G-protein complex; receptor complex]

Mayya2005_STATmodule: BIOMD0000000167v0.0.1

The model reproduces Fig 2B of the paper. Model successfully reproduced using MathSBML. To the extent possible under la…

Details

Signal transducer and actuator of transcription (STATs) are a family of transcription factors activated by various cytokines, growth factors and hormones. They are important mediators of immune responses and growth and differentiation of various cell types. The STAT signalling system represents a defined functional module with a pattern of signalling that is conserved from flies to mammals. In order to probe and gain insights into the signalling properties of the STAT module by computational means, we developed a simple non-linear ordinary differential equations model within the 'Virtual Cell' framework. Our results demonstrate that the STAT module can operate as a 'biphasic amplitude filter' with an ability to amplify input signals within a specific intermediate range. We show that dimerisation of phosphorylated STAT is crucial for signal amplification and the amplitude filtering function. We also demonstrate that maximal amplification at intermediate levels of STAT activation is a moderately robust property of STAT module. We propose that these observations can be extrapolated to the analogous SMAD signalling module. link: http://identifiers.org/pubmed/17091582

Parameters:

NameDescription
stat_expMax=-0.06 μmol*l^(-1)*μm^(-2)*s^(-1); Ks_exp=0.6 μmol*l^(-1)Reaction: stat_sol => stat_nuc, Rate Law: nuc*stat_expMax*stat_nuc*1/(Ks_exp+stat_nuc)*nm
stat_impMax=0.003 μmol*l^(-1)*μm^(-2)*s^(-1); Ks_imp=3.0 μmol*l^(-1)Reaction: stat_sol => stat_nuc, Rate Law: nuc*stat_impMax*stat_sol*1/(Ks_imp+stat_sol)*nm
Kf_PstatDimerisation=0.6 μmol^(-1)*l*s^(-1); Kr_PstatDimerisation=0.03 s^(-1)Reaction: Pstat_sol => PstatDimer_sol, Rate Law: (Kf_PstatDimerisation*Pstat_sol^2+(-Kr_PstatDimerisation*PstatDimer_sol))*sol
PstatDimer_impMax=0.045 μmol*l^(-1)*μm^(-2)*s^(-1); Kpsd_imp=0.3 μmol*l^(-1)Reaction: PstatDimer_sol => PstatDimer_nuc, Rate Law: PstatDimer_impMax*PstatDimer_sol*1/(Kpsd_imp+PstatDimer_sol)*nm
Kcat_phos=1.0 s^(-1); Km_phos=4.0 μmol*l^(-1)Reaction: stat_sol => Pstat_sol + species_test; statKinase_sol, Rate Law: Kcat_phos*statKinase_sol*stat_sol*1/(Km_phos+stat_sol)*sol
Km_dephos=2.0 μmol*l^(-1); Kcat_dephos=1.0 s^(-1)Reaction: Pstat_nuc => stat_nuc; statPhosphatase_nuc, Rate Law: Kcat_dephos*statPhosphatase_nuc*Pstat_nuc*1/(Km_dephos+Pstat_nuc)*nuc

States:

NameDescription
Pstat sol[Signal transducer and activator of transcription 1-alpha/beta]
PstatDimer nuc[Signal transducer and activator of transcription 1-alpha/beta]
statKinase solstatKinase_sol
species testspecies_test
PstatDimer sol[Signal transducer and activator of transcription 1-alpha/beta]
Pstat nuc[Signal transducer and activator of transcription 1-alpha/beta]
stat sol[Signal transducer and activator of transcription 1-alpha/beta]
stat nuc[Signal transducer and activator of transcription 1-alpha/beta]

Mazet2020 - model of the PI cycle: MODEL2006300001v0.0.1

<notes xmlns="http://www.sbml.org/sbml/level2/version4"> <body xmlns="http://www.w3.org/1…

Details

The phosphatidylinositol (PI) cycle is central to eukaryotic cell signaling. Its complexity, due to the number of reactions and lipid and inositol phosphate intermediates involved makes it difficult to analyze experimentally. Computational modelling approaches are seen as a way forward to elucidate complex biological regulatory mechanisms when this cannot be achieved solely through experimental approaches. Whilst mathematical modelling is well established in informing biological systems, many models are often informed by data sourced from different cell types (mosaic data), to inform model parameters. For instance, kinetic rate constants are often determined from purified enzyme data in vitro or use experimental concentrations obtained from multiple unrelated cell types. Thus they do not represent any specific cell type nor fully capture cell specific behaviours. In this work, we develop a model of the PI cycle informed by in-vivo omics data taken from a single cell type, namely platelets. Our model recapitulates the known experimental dynamics before and after stimulation with different agonists and demonstrates the importance of lipid- and protein-binding proteins in regulating second messenger outputs. Furthermore, we were able to make a number of predictions regarding the regulation of PI cycle enzymes and the importance of the number of receptors required for successful GPCR signaling. We then consider how pathway behavior differs, when fully informed by data for HeLa cells and show that model predictions remain relatively consistent. However, when informed by mosaic experimental data model predictions greatly vary. Our work illustrates the risks of using mosaic datasets from unrelated cell types which leads to over 75% of outputs not fitting with expected behaviors. link:

Mazumdar2008 - Genome-scale metabolic network of Porphyromonas gingivalis (iVM679): MODEL1507180038v0.0.1

Mazumdar2008 - Genome-scale metabolic network of Porphyromonas gingivalis (iVM679)This model is described in the article…

Details

The microbial community present in the human mouth is engaged in a complex network of diverse metabolic activities. In addition to serving as energy and building-block sources, metabolites are key players in interspecies and host-pathogen interactions. Metabolites are also implicated in triggering the local inflammatory response, which can affect systemic conditions such as atherosclerosis, obesity, and diabetes. While the genome of several oral pathogens has been sequenced, quantitative understanding of the metabolic functions of any oral pathogen at the system level has not been explored yet. Here we pursue the computational construction and analysis of the genome-scale metabolic network of Porphyromonas gingivalis, a gram-negative anaerobe that is endemic in the human population and largely responsible for adult periodontitis. Integrating information from the genome, online databases, and literature screening, we built a stoichiometric model that encompasses 679 metabolic reactions. By using flux balance approaches and automated network visualization, we analyze the growth capacity under amino-acid-rich medium and provide evidence that amino acid preference and cytotoxic by-product secretion rates are suitably reproduced by the model. To provide further insight into the basic metabolic functions of P. gingivalis and suggest potential drug targets, we study systematically how the network responds to any reaction knockout. We focus specifically on the lipopolysaccharide biosynthesis pathway and identify eight putative targets, one of which has been recently verified experimentally. The current model, which is amenable to further experimental testing and refinements, could prove useful in evaluating the oral microbiome dynamics and in the development of novel biomedical applications. link: http://identifiers.org/pubmed/18931137

Mbodj2016 - Mesoderm specification during Drosophila development: MODEL1607310000v0.0.1

Mbodj2016 - Mesoderm specification during Drosophila developmentThis model is described in the article: [Qualitative Dy…

Details

Given the complexity of developmental networks, it is often difficult to predict the effect of genetic perturbations, even within coding genes. Regulatory factors generally have pleiotropic effects, exhibit partially redundant roles, and regulate highly interconnected pathways with ample cross-talk. Here, we delineate a logical model encompassing 48 components and 82 regulatory interactions involved in mesoderm specification during Drosophila development, thereby providing a formal integration of all available genetic information from the literature. The four main tissues derived from mesoderm correspond to alternative stable states. We demonstrate that the model can predict known mutant phenotypes and use it to systematically predict the effects of over 300 new, often non-intuitive, loss- and gain-of-function mutations, and combinations thereof. We further validated several novel predictions experimentally, thereby demonstrating the robustness of model. Logical modelling can thus contribute to formally explain and predict regulatory outcomes underlying cell fate decisions. link: http://identifiers.org/pubmed/27599298

McAllister1975_CardiacPurkinjeFibres: MODEL9808533471v0.0.1

This a model from the article: Reconstruction of the electrical activity of cardiac Purkinje fibres. McAllister RE,…

Details

  1. The electrical activity of Cardiac Purkinje fibres was reconstructed using a mathematical model of the membrane current. The individual components of ionic curent were described by equations which wee based as closely as possible on previous experiments using the voltage clamp technique. 2. Membrane action potentials and pace-maker activity were calculated and compared with time course of underlying changes in two functionally distinct outeard currents, iX1 and iK2. 3. The repolarization of the theoretical action potential is triggered by the onset of iX1, which becomes activated over the plateau range of potentials. iK2 also activates during the plateau but does not play a controlling role in the repolarization. Hwever, iK2 does govern the slow pace-maker depolarization through its subsequent deactivation at negative potentials. 4. The individual phases of the calculated action potential and their 'experimental' modifications were compared with published records. The upstroke is generated by a Hodgkin-Huxley type sodium conductance (gNa), and rises with a maximum rate of 478 V/sec, somewhat less than experimentally observed values ( up to 800 V/sec). The discrepancy is discussed in relation to experimental attempts at measuring gNa. 5. The ole of the transient outward chloride current (called igr) was studied in calculations of the rapid phase of repolarization and 'notch' configuration...

link: http://identifiers.org/pubmed/1185607

McAuley2012 - Whole-body Cholesterol Metabolism: BIOMD0000000434v0.0.1

McAuley2012 - Whole-body Cholesterol MetabolismLipid metabolism has a key role to play in human longevity and healthy ag…

Details

BACKGROUND: Global demographic changes have stimulated marked interest in the process of aging. There has been, and will continue to be, an unrelenting rise in the number of the oldest old ( >85 years of age). Together with an ageing population there comes an increase in the prevalence of age related disease. Of the diseases of ageing, cardiovascular disease (CVD) has by far the highest prevalence. It is regarded that a finely tuned lipid profile may help to prevent CVD as there is a long established relationship between alterations to lipid metabolism and CVD risk. In fact elevated plasma cholesterol, particularly Low Density Lipoprotein Cholesterol (LDL-C) has consistently stood out as a risk factor for having a cardiovascular event. Moreover it is widely acknowledged that LDL-C may rise with age in both sexes in a wide variety of groups. The aim of this work was to use a whole-body mathematical model to investigate why LDL-C rises with age, and to test the hypothesis that mechanistic changes to cholesterol absorption and LDL-C removal from the plasma are responsible for the rise. The whole-body mechanistic nature of the model differs from previous models of cholesterol metabolism which have either focused on intracellular cholesterol homeostasis or have concentrated on an isolated area of lipoprotein dynamics. The model integrates both current and previously published data relating to molecular biology, physiology, ageing and nutrition in an integrated fashion. RESULTS: The model was used to test the hypothesis that alterations to the rate of cholesterol absorption and changes to the rate of removal of LDL-C from the plasma are integral to understanding why LDL-C rises with age. The model demonstrates that increasing the rate of intestinal cholesterol absorption from 50% to 80% by age 65 years can result in an increase of LDL-C by as much as 34 mg/dL in a hypothetical male subject. The model also shows that decreasing the rate of hepatic clearance of LDL-C gradually to 50% by age 65 years can result in an increase of LDL-C by as much as 116 mg/dL. CONCLUSIONS: Our model clearly demonstrates that of the two putative mechanisms that have been implicated in the dysregulation of cholesterol metabolism with age, alterations to the removal rate of plasma LDL-C has the most significant impact on cholesterol metabolism and small changes to the number of hepatic LDL receptors can result in a significant rise in LDL-C. This first whole-body systems based model of cholesterol balance could potentially be used as a tool to further improve our understanding of whole-body cholesterol metabolism and its dysregulation with age. Furthermore, given further fine tuning the model may help to investigate potential dietary and lifestyle regimes that have the potential to mitigate the effects aging has on cholesterol metabolism. link: http://identifiers.org/pubmed/23046614

Parameters:

NameDescription
k18=0.068Reaction: species_23 => species_7; species_18, species_23, species_18, Rate Law: k18*species_23*species_18
khrs=100.0Reaction: species_19 => species_18; species_19, species_7, species_19, species_7, Rate Law: khrs*species_19/species_7
k17=0.38Reaction: species_21 => species_23; species_24, species_21, species_24, Rate Law: k17*species_21*species_24
k5=2.66Reaction: species_7 => species_4; species_4, species_7, species_4, Rate Law: k5*species_7/species_4
kprs=100.0Reaction: species_26 => species_25; species_11, species_26, species_11, Rate Law: kprs*species_26/species_11
k1=5.0E-6Reaction: species_23 => species_11; species_23, Rate Law: k1*species_23
k1=6.0Reaction: species_4 => species_5; species_4, Rate Law: k1*species_4
k8=5.0E-4Reaction: species_9 => species_10; species_11, species_11, Rate Law: k8*species_11
k1=0.016Reaction: species_7 => species_17; species_7, Rate Law: k1*species_7
k1=0.01Reaction: species_18 => species_20; species_18, Rate Law: k1*species_18
k6=5.286E-4Reaction: species_2 => species_7; species_5, species_2, species_5, Rate Law: k6*species_2*species_5
k9=1.0Reaction: species_7 => species_13; species_14, species_7, species_14, species_7, Rate Law: k9*species_14*species_7
k15=0.43Reaction: species_17 => species_21; species_17, species_22, species_17, species_22, Rate Law: k15*species_17*species_22
k26=1.5E-5Reaction: species_11 + species_10 => species_30; species_31, species_11, species_10, species_31, Rate Law: k26*species_11*species_10*species_31
k27=0.01Reaction: species_30 => species_17; species_33, species_30, species_33, Rate Law: k27*species_30*species_33
BS=5.0; BCRt=55326.0; BCRmax=2000.0Reaction: species_7 => species_2; species_7, species_7, Rate Law: BCRmax/(1+(BCRt/species_7)^BS)
k29=0.05Reaction: species_30 => species_7; species_34, species_30, species_34, Rate Law: k29*species_30*species_34
HCSt=93925.0; HCSmax=500.0; HS=5.0Reaction: species_12 => species_7; species_7, species_7, Rate Law: HCSmax/(1+(species_7/HCSt)^HS)
k1=1.0Reaction: species_1 => species_2; species_1, Rate Law: k1*species_1
k1=5.0E-4Reaction: species_11 => species_29; species_11, Rate Law: k1*species_11
k1=0.054Reaction: species_21 => species_7; species_21, Rate Law: k1*species_21
k28=0.001Reaction: species_30 => species_23; species_33, species_30, species_33, Rate Law: k28*species_30*species_33
k1=4.29Reaction: species_5 => species_4; species_5, Rate Law: k1*species_5
k10=5.998Reaction: species_13 => species_7; species_15, species_13, species_15, species_13, Rate Law: k10*species_15*species_13
k1=0.0496Reaction: species_17 => species_7; species_17, Rate Law: k1*species_17
k1=0.005Reaction: species_23 => species_7; species_23, Rate Law: k1*species_23
k24=0.1068Reaction: species_28 => species_11; species_15, species_15, species_28, Rate Law: k24*species_15*species_28
ICSmax=100.0; IS=5.0; ICt=3120.0Reaction: species_3 => species_2; species_2, species_2, Rate Law: ICSmax/(1+(species_2/ICt)^IS)
k1=0.856Reaction: species_5 => species_6; species_5, Rate Law: k1*species_5
k11=0.005Reaction: species_16 => species_10; species_11, species_11, Rate Law: k11*species_11
k20=0.00675Reaction: species_23 => species_11; species_25, species_25, species_23, Rate Law: k20*species_25*species_23
k7=5.286E-4Reaction: species_2 => species_8; species_5, species_2, species_5, Rate Law: k7*species_2*species_5
k23=0.017386Reaction: species_11 => species_28; species_14, species_14, species_11, Rate Law: k23*species_14*species_11
PCSS=5.0; PPCt=80342.0; PCSmax=500.0Reaction: species_32 => species_11; species_11, Rate Law: PCSmax/(1+(species_11/PPCt)^PCSS)

States:

NameDescription
species 9INHDLS
species 27PLDLRD
species 1[cholesterol]
species 18[Low-density lipoprotein receptor]
species 4[bile salt]
species 16HNHDLS
species 20HLDLRD
species 28[cholesteryl ester]
species 25[Low-density lipoprotein receptor]
species 21[Low-density lipoprotein receptor]
species 17[Very low-density lipoprotein receptor]
species 29PSS
species 30[Vigilin]
species 5[bile salt]
species 8[cholesterol]
species 32PCS
species 12HCS
species 2[cholesterol]
species 6[bile salt]
species 19HLDLRsS
species 10[Vigilin]
species 11[cholesterol]
species 3ICS
species 23[Low-density lipoprotein receptor]
species 7[cholesterol]
species 26PLDLRsS
species 13[cholesteryl ester]

McClean2007_CrossTalk: BIOMD0000000116v0.0.1

This model encoded according to the paper *Cross-talk and decision making in MAP kinase pathways.* Supplementary Figure…

Details

Cells must respond specifically to different environmental stimuli in order to survive. The signal transduction pathways involved in sensing these stimuli often share the same or homologous proteins. Despite potential cross-wiring, cells show specificity of response. We show, through modeling, that the physiological response of such pathways exposed to simultaneous and temporally ordered inputs can demonstrate system-level mechanisms by which pathways achieve specificity. We apply these results to the hyperosmolar and pheromone mitogen-activated protein (MAP) kinase pathways in the yeast Saccharomyces cerevisiae. These two pathways specifically sense osmolar and pheromone signals, despite sharing a MAPKKK, Ste11, and having homologous MAPKs (Fus3 and Hog1). We show that in a single cell, the pathways are bistable over a range of inputs, and the cell responds to only one stimulus even when exposed to both. Our results imply that these pathways achieve specificity by filtering out spurious cross-talk through mutual inhibition. The variability between cells allows for heterogeneity of the decisions. link: http://identifiers.org/pubmed/17259986

Parameters:

NameDescription
parameter_0 = 10.0; parameter_7 = 8.5; parameter_8 = 1.0; parameter_6 = 1.0Reaction: => species_0, Rate Law: compartment_0*parameter_6*parameter_7/(1+parameter_7/parameter_8)*(parameter_0-species_0)
parameter_10 = 1.0; parameter_2 = 10.0Reaction: species_1 => species_2, Rate Law: compartment_0*parameter_10*species_1*(parameter_2-species_2)
parameter_12 = 0.0; parameter_4 = 10.0Reaction: species_0 => species_4, Rate Law: compartment_0*parameter_12*species_0*(parameter_4-species_4)
parameter_9 = 1.0; parameter_1 = 10.0Reaction: species_0 => species_1, Rate Law: compartment_0*parameter_9*species_0*(parameter_1-species_1)
parameter_16 = 1.0; parameter_14 = 5.0; parameter_3 = 10.0; parameter_15 = 1.0Reaction: => species_3, Rate Law: compartment_0*parameter_15*parameter_14/(1+parameter_14/parameter_16)*(parameter_3-species_3)
parameter_4 = 10.0; parameter_17 = 1.0Reaction: species_3 => species_4, Rate Law: compartment_0*parameter_17*species_3*(parameter_4-species_4)
parameter_18 = 1.0; parameter_5 = 10.0Reaction: species_4 => species_5, Rate Law: compartment_0*parameter_18*species_4*(parameter_5-species_5)
parameter_19 = 1.0; parameter_11 = 1.0Reaction: species_5 => ; species_2, Rate Law: compartment_0*parameter_11*species_5*species_2/(1+species_5/parameter_19)
parameter_12 = 0.0; parameter_1 = 10.0Reaction: species_3 => species_1, Rate Law: compartment_0*parameter_12*species_3*(parameter_1-species_1)
parameter_13 = 1.0; parameter_11 = 1.0Reaction: species_2 => ; species_5, Rate Law: compartment_0*parameter_11*species_5*species_2/(1+species_2/parameter_13)

States:

NameDescription
species 2[Mitogen-activated protein kinase FUS3]
species 3[Serine/threonine-protein kinase STE11]
species 0[Serine/threonine-protein kinase STE11]
species 1[Serine/threonine-protein kinase STE7]
species 4[MAP kinase kinase PBS2]
species 5[Mitogen-activated protein kinase HOG1]

McDougal2017 - Metabolism in ischemic cardiomyocytes: BIOMD0000000961v0.0.1

Heart disease remains the leading cause of death globally. Although reperfusion following myocardial ischemia can preven…

Details

Heart disease remains the leading cause of death globally. Although reperfusion following myocardial ischemia can prevent death by restoring nutrient flow, ischemia/reperfusion injury can cause significant heart damage. The mechanisms that drive ischemia/reperfusion injury are not well understood; currently, few methods can predict the state of the cardiac muscle cell and its metabolic conditions during ischemia. Here, we explored the energetic sustainability of cardiomyocytes, using a model for cellular metabolism to predict the levels of ATP following hypoxia. We modeled glycolytic metabolism with a system of coupled ordinary differential equations describing the individual metabolic reactions within the cardiomyocyte over time. Reduced oxygen levels and ATP consumption rates were simulated to characterize metabolite responses to ischemia. By tracking biochemical species within the cell, our model enables prediction of the cell's condition up to the moment of reperfusion. The simulations revealed a distinct transition between energetically sustainable and unsustainable ATP concentrations for various energetic demands. Our model illustrates how even low oxygen concentrations allow the cell to perform essential functions. We found that the oxygen level required for a sustainable level of ATP increases roughly linearly with the ATP consumption rate. An extracellular O2 concentration of ∼0.007 mm could supply basic energy needs in non-beating cardiomyocytes, suggesting that increased collateral circulation may provide an important source of oxygen to sustain the cardiomyocyte during extended ischemia. Our model provides a time-dependent framework for studying various intervention strategies to change the outcome of reperfusion. link: http://identifiers.org/pubmed/28487363

McLean1991 - Behaviour of HIV in the presence of zidovudine: BIOMD0000000967v0.0.1

A new mechanism is proposed for the apparent breakthrough of HIV that occurs approximately 6 months after the commenceme…

Details

A new mechanism is proposed for the apparent breakthrough of HIV that occurs approximately 6 months after the commencement of therapy with zidovudine (AZT). Using a simple mathematical model of the interacting population dynamics of HIV and its major host cell in the circulation (the CD4+ lymphocyte), predicted patterns of HIV plasma viraemia in the weeks following treatment with zidovudine are generated. These are in close agreement with observed patterns despite the fact that the model contains no mechanisms for the development of drug-resistant strains of virus. It is suggested that the patterns of viral abundance observed during the first 6 months after treatment may be the result of non-linearities in the interactions between HIV and CD4+ cells, and that it is only after the first post-treatment burst of viral production that drug resistance plays an important role. link: http://identifiers.org/pubmed/1677807

Mears1997_CRAC_PancreaticBetaCells: BIOMD0000000375v0.0.1

This a model from the article: Evidence that calcium release-activated current mediates the biphasic electrical acti…

Details

The electrical response of pancreatic beta-cells to step increases in glucose concentration is biphasic, consisting of a prolonged depolarization with action potentials (Phase 1) followed by membrane potential oscillations known as bursts. We have proposed that the Phase 1 response results from the combined depolarizing influences of potassium channel closure and an inward, nonselective cation current (ICRAN) that activates as intracellular calcium stores empty during exposure to basal glucose (Bertram et al., 1995). The stores refill during Phase 1, deactivating ICRAN and allowing steady-state bursting to commence. We support this hypothesis with additional simulations and experimental results indicating that Phase 1 duration is sensitive to the filling state of intracellular calcium stores. First, the duration of the Phase 1 transient increases with duration of prior exposure to basal (2.8 mM) glucose, reflecting the increased time required to fill calcium stores that have been emptying for longer periods. Second, Phase 1 duration is reduced when islets are exposed to elevated K+ to refill calcium stores in the presence of basal glucose. Third, when extracellular calcium is removed during the basal glucose exposure to reduce calcium influx into the stores, Phase 1 duration increases. Finally, no Phase 1 is observed following hyperpolarization of the beta-cell membrane with diazoxide in the continued presence of 11 mm glucose, a condition in which intracellular calcium stores remain full. Application of carbachol to empty calcium stores during basal glucose exposure did not increase Phase 1 duration as the model predicts. Despite this discrepancy, the good agreement between most of the experimental results and the model predictions provides evidence that a calcium release-activated current mediates the Phase 1 electrical response of the pancreatic beta-cell. link: http://identifiers.org/pubmed/9002424

Parameters:

NameDescription
J_mem_tot = -2.34898089778648E-5; lambda_er = 250.0; J_er_tot = 0.0359076237623762Reaction: Ca_i = J_er_tot/lambda_er+J_mem_tot, Rate Law: J_er_tot/lambda_er+J_mem_tot
sigma_er = 1.0; lambda_er = 250.0; J_er_tot = 0.0359076237623762Reaction: Ca_er_Ca_equations = (-J_er_tot)/(lambda_er*sigma_er), Rate Law: (-J_er_tot)/(lambda_er*sigma_er)
tau_n = 9.085746273364; lambda_n = 1.85; n_infinity = 4.67956725632935E-4Reaction: n = lambda_n*(n_infinity-n)/tau_n, Rate Law: lambda_n*(n_infinity-n)/tau_n
i_K_Ca = 3.45489443378119; i_leak = 0.0; i_CRAC = -5.81489940359721; Cm = 6158.0; i_Ca = -1342.58335216182; i_K = 17.55; i_K_ATP = 1350.0Reaction: V_membrane = (-(i_Ca+i_K+i_K_ATP+i_K_Ca+i_CRAC+i_leak))/Cm, Rate Law: (-(i_Ca+i_K+i_K_ATP+i_K_Ca+i_CRAC+i_leak))/Cm
jm_infinity = 0.0179862099620915; tau_j = 8145.05572085199Reaction: jm = (jm_infinity-jm)/tau_j, Rate Law: (jm_infinity-jm)/tau_j

States:

NameDescription
Ca i[calcium(2+)]
V membrane[membrane potential]
jmjm
Ca er Ca equations[calcium(2+)]
n[delayed rectifier potassium channel activity]

Mellor2012_LipooxygenasePathway: BIOMD0000000415v0.0.1

This model is from the article: Reduction of off-flavor generation in soybean homogenates: a mathematical model. M…

Details

The generation of off-flavors in soybean homogenates such as n-hexanal via the lipoxygenase (LOX) pathway can be a problem in the processed food industry. Previous studies have examined the effect of using soybean varieties missing one or more of the 3 LOX isozymes on n-hexanal generation. A dynamic mathematical model of the soybean LOX pathway using ordinary differential equations was constructed using parameters estimated from existing data with the aim of predicting how n-hexanal generation could be reduced. Time-course simulations of LOX-null beans were run and compared with experimental results. Model L(2), L(3), and L(12) beans were within the range relative to the wild type found experimentally, with L(13) and L(23) beans close to the experimental range. Model L(1) beans produced much more n-hexanal relative to the wild type than those in experiments. Sensitivity analysis indicates that reducing the estimated K(m) parameter for LOX isozyme 3 (L-3) would improve the fit between model predictions and experimental results found in the literature. The model also predicts that increasing L-3 or reducing L-2 levels within beans may reduce n-hexanal generation.This work describes the use of mathematics to attempt to quantify the enzyme-catalyzed conversions of compounds in soybean homogenates into undesirable flavors, primarily from the compound n-hexanal. The effect of different soybean genotypes and enzyme kinetic constants was also studied, leading to recommendations on which combinations might minimize off-flavor levels and what further work might be carried out to substantiate these conclusions. link: http://identifiers.org/pubmed/21535565

Parameters:

NameDescription
parameter_7 = 0.05; parameter_9 = 0.038475Reaction: species_8 => species_15, Rate Law: compartment_1*parameter_9*species_8/(parameter_7+species_8)
parameter_5 = 0.49; parameter_6 = 0.00255Reaction: species_1 => species_7 + species_8 + species_9 + species_10 + species_11 + species_12 + species_13 + species_14, Rate Law: compartment_1*parameter_6*species_1/(parameter_5+species_1)
parameter_4 = 0.039; parameter_3 = 0.49Reaction: species_1 => species_7 + species_8 + species_9 + species_10 + species_11 + species_12 + species_13 + species_14, Rate Law: compartment_1*parameter_4*species_1/(parameter_3+species_1)
parameter_1 = 0.49; parameter_2 = 0.00825Reaction: species_1 => species_7 + species_8 + species_9 + species_10 + species_11 + species_12 + species_13 + species_14, Rate Law: compartment_1*parameter_2*species_1/(parameter_1+species_1)
parameter_7 = 0.05; parameter_8 = 0.285Reaction: species_7 => species_15, Rate Law: compartment_1*parameter_8*species_7/(parameter_7+species_7)

States:

NameDescription
species 14[hydroperoxide]
species 9[hydroperoxide]
species 10[hydroperoxide]
species 11[hydroperoxide]
species 1[linoleic acid]
species 8[hydroperoxide]
species 12[hydroperoxide]
species 7[hydroperoxide]
species 15[6184]
species 13[hydroperoxide]

Merola2008 - An insight into tumor dormancy equilibrium via the analysis of its domain of attraction: BIOMD0000000911v0.0.1

An insight into tumor dormancy equilibrium via the analysis of its domain of attraction A. Merola, C. Cosentino *, F. Am…

Details

A B S T R A C T The trajectories of the dynamic system which regulates the competition between the populations of malignant cells and immune cells may tend to an asymptotically stable equilibrium in which the sizes of these populations do not vary, which is called tumor dormancy. Especially for lower steady-state sizes of the population of malignant cells, this equilibrium represents a desirable clinical condition since the tumor growth is blocked. In this context, it is of mandatory importance to analyze the robustness of this clinical favorable state of health in the face of perturbations. To this end, the paper presents an optimization technique to determine whether an assigned rectangular region, which surrounds an asymptotically stable equilibrium point of a quadratic systems, is included into the domain of attraction of the equilibrium itself. The biological relevance of the application of this technique to the analysis of tumor growth dynamics is shown on the basis of a recent quadraticmodel of the tumor–immune system competition dynamics. Indeed the application of the proposedmethodology allows to ensure that a given safety region, determined on the basis of clinical considerations, belongs to the domain of attraction of the tumor blocked equilibrium; therefore for the set of perturbed initial conditions which belong to such region, the convergence to the healthy steady state is guaranteed. The proposed methodology can also provide an optimal strategy for cancer treatment. link: http://identifiers.org/doi/10.1016/j.bspc.2008.02.001

Parameters:

NameDescription
r = 0.9; k1 = 0.8; q = 10.0Reaction: => M, Rate Law: compartment*(q+r*M*(1-M/k1))
alpha = 0.3Reaction: M => ; N, Rate Law: compartment*alpha*M*N
beta = 0.1; d2 = 0.03Reaction: Z => ; N, Rate Law: compartment*(beta*N*Z+d2*Z)
beta = 0.1Reaction: => N; Z, Rate Law: compartment*beta*N*Z
k2 = 0.7; s = 0.8Reaction: => Z, Rate Law: compartment*s*Z*(1-Z/k2)
d1 = 0.02Reaction: N =>, Rate Law: compartment*d1*N

States:

NameDescription
ZZ
M[Neoplastic Cell]
NN

Messiha2013 - combined glycolysis and pentose phosphate pathway model: BIOMD0000000503v0.0.1

Messiha2013 - combined glycolysis and pentose phosphate pathway model[BIOMD0000000502](http://identifiers.org/biomodels.…

Details

We present the quantification and kinetic characterisation of the enzymes of the pentose phosphate pathway in Saccharomyces cerevisiae. The data are combined into a mathematical model that describes the dynamics of this system and allows for the predicting changes in metabolite concentrations and fluxes in response to perturbations. We use the model to study the response of yeast to a glucose pulse. We then combine the model with an existing glycolysis one to study the effect of oxidative stress on carbohydrate metabolism. The combination of these two models was made possible by the standardized enzyme kinetic experiments carried out in both studies. This work demonstrates the feasibility of constructing larger network models by merging smaller pathway models. link: http://identifiers.org/doi/10.7287/peerj.preprints.146v2

Parameters:

NameDescription
sum_NAD = 1.59 mMReaction: NADH = sum_NAD-NAD, Rate Law: missing
sum_UxP = 1.39784619487425 mMReaction: UDG = (sum_UxP-UTP)-UDP, Rate Law: missing
Kudg=0.886 mM; Vmax=0.49356 mM per s; Kg6p=3.8 mMReaction: G6P + UDG => T6P + UDP; TPS1, TPS2, G6P, UDG, Rate Law: cell*Vmax*G6P*UDG/(Kg6p*Kudg)/((1+G6P/Kg6p)*(1+UDG/Kudg))
k=1.0 per sReaction: E4P => ; E4P, Rate Law: cell*k*E4P
Vmax=0.12762 mM per s; Kg1p=0.023 mM; Kg6p=0.05 mM; Keq=0.1667 dimensionlessReaction: G6P => G1P; PGM1, PGM2, G6P, G1P, Rate Law: cell*Vmax*(G6P/Kg6p-G1P/(Kg6p*Keq))/(1+G6P/Kg6p+G1P/Kg1p)
Kg6p_GLK1=30.0 mM; Kg6p_HXK1=30.0 mM; kcat_HXK1=10.2 per s; Kg6p_HXK2=30.0 mM; Kit6p_HXK1=0.2 mM; Katp_GLK1=0.865 mM; Katp_HXK1=0.293 mM; kcat_GLK1=0.0721 per s; Keq=2000.0 dimensionless; Katp_HXK2=0.195 mM; kcat_HXK2=63.1 per s; Kglc_HXK2=0.2 mM; Kit6p_HXK2=0.04 mM; Kglc_GLK1=0.0106 mM; Kglc_HXK1=0.15 mM; Kadp_HXK1=0.23 mM; Kadp_GLK1=0.23 mM; Kadp_HXK2=0.23 mMReaction: GLC + ATP => G6P + ADP; HXK1, T6P, HXK2, GLK1, HXK1, GLC, ATP, G6P, ADP, T6P, HXK2, GLK1, Rate Law: cell*HXK1*kcat_HXK1*(GLC*ATP/(Kglc_HXK1*Katp_HXK1)-G6P*ADP/(Kglc_HXK1*Katp_HXK1*Keq))/((1+GLC/Kglc_HXK1+G6P/Kg6p_HXK1+T6P/Kit6p_HXK1)*(1+ATP/Katp_HXK1+ADP/Kadp_HXK1))+cell*HXK2*kcat_HXK2*(GLC*ATP/(Kglc_HXK2*Katp_HXK2)-G6P*ADP/(Kglc_HXK2*Katp_HXK2*Keq))/((1+GLC/Kglc_HXK2+G6P/Kg6p_HXK2+T6P/Kit6p_HXK2)*(1+ATP/Katp_HXK2+ADP/Kadp_HXK2))+cell*GLK1*kcat_GLK1*(GLC*ATP/(Kglc_GLK1*Katp_GLK1)-G6P*ADP/(Kglc_GLK1*Katp_GLK1*Keq))/((1+GLC/Kglc_GLK1+G6P/Kg6p_GLK1)*(1+ATP/Katp_GLK1+ADP/Kadp_GLK1))
Vmax=6.16 mM per s; Katp=3.0 mMReaction: ATP => ADP; ATP, Rate Law: cell*Vmax*ATP/Katp/(1+ATP/Katp)
Kadp=0.2 mM; Kbpg=0.003 mM; kcat=58.6 per s; Katp=1.99 mM; Kp3g=4.58 mM; nHadp=2.0 dimensionless; Keq=3200.0 dimensionlessReaction: ADP + BPG => ATP + P3G; PGK1, PGK1, ADP, BPG, P3G, ATP, Rate Law: cell*PGK1*kcat*(ADP/Kadp)^(nHadp-1)*(BPG*ADP/(Kbpg*Kadp)-P3G*ATP/(Kbpg*Kadp*Keq))/((1+BPG/Kbpg+P3G/Kp3g)*(1+(ADP/Kadp)^nHadp+ATP/Katp))
Kx5p=7.7 mM; Keq=1.4 dimensionless; kcat=4020.0 per s; Kru5p=5.97 mMReaction: Ru5P => X5P; RPE1, RPE1, Ru5P, X5P, Rate Law: cell*RPE1*kcat*(Ru5P-X5P/Keq)/Kru5p/(1+Ru5P/Kru5p+X5P/Kx5p)
Knadh=0.023 mM; Katp=0.73 mM; Kg3p=1.2 mM; Kfbp=4.8 mM; Keq=10000.0 dimensionless; Kadp=2.0 mM; Vmax=0.783333333333333 mM per s; Kdhap=0.54 mM; Knad=0.93 mMReaction: DHAP + NADH => G3P + NAD; ADP, ATP, F16bP, GPD1, GPD2, DHAP, NADH, G3P, NAD, F16bP, ATP, ADP, Rate Law: cell*Vmax/(Kdhap*Knadh)*(DHAP*NADH-G3P*NAD/Keq)/((1+F16bP/Kfbp+ATP/Katp+ADP/Kadp)*(1+DHAP/Kdhap+G3P/Kg3p)*(1+NADH/Knadh+NAD/Knad))
k=0.00554339592436782 per_mM_per_sReaction: AcAld + NAD => ACE + NADH; AcAld, NAD, Rate Law: cell*k*AcAld*NAD
Keq=0.19 dimensionless; Kp2g=1.41 mM; kcat=400.0 per s; Kp3g=1.2 mMReaction: P3G => P2G; GPM1, GPM1, P3G, P2G, Rate Law: cell*GPM1*kcat*(P3G/Kp3g-P2G/(Kp3g*Keq))/(1+P3G/Kp3g+P2G/Kp2g)
sum_AxP = 6.02 mMReaction: AMP = (sum_AxP-ATP)-ADP, Rate Law: missing
k=0.0745258294103764 per_mM_per_sReaction: UDP + ATP => UTP + ADP; UDP, ATP, Rate Law: cell*k*UDP*ATP
Knadp=0.045 mM; kcat=189.0 per s; Kg6p=0.042 mM; Knadph=0.017 mM; Kg6l=0.01 mMReaction: G6P + NADP => G6L + NADPH; ZWF1, ZWF1, G6P, NADP, G6L, NADPH, Rate Law: cell*ZWF1*kcat*G6P*NADP/(Kg6p*Knadp)/((1+G6P/Kg6p+G6L/Kg6l)*(1+NADP/Knadp+NADPH/Knadph))
Kru5p=2.47 mM; kcat=335.0 per s; Kiru5p=9.88 mM; Keq=4.0 dimensionless; Kr5p=5.7 mMReaction: Ru5P => R5P; RKI1, RKI1, Ru5P, R5P, Rate Law: cell*RKI1*kcat*(Ru5P-R5P/Keq)/Kru5p/(1+Ru5P/Kru5p+R5P/Kr5p)
Kg6l=0.8 mM; kcat=4.3 per s; Kp6g=0.5 mMReaction: G6L => P6G; SOL3, SOL3, G6L, P6G, Rate Law: cell*SOL3*kcat*G6L/Kg6l/(1+G6L/Kg6l+P6G/Kp6g)
sum_NADP = 0.33 mMReaction: NADP = sum_NADP-NADPH, Rate Law: missing
Kgap_TAL1=0.272 mM; Ke4p_NQM1=0.305 mM; kcat_NQM1=0.694 per s; Kf6p_TAL1=1.44 mM; kcat_TAL1=0.694 per s; Kgap_NQM1=0.272 mM; Ks7p_NQM1=0.786 mM; Kf6p_NQM1=1.04 mM; Ke4p_TAL1=0.362 mM; Keq=1.05 dimensionless; Ks7p_TAL1=0.786 mMReaction: GAP + S7P => F6P + E4P; TAL1, NQM1, TAL1, GAP, S7P, F6P, E4P, NQM1, Rate Law: cell*(TAL1*kcat_TAL1*(GAP*S7P-F6P*E4P/Keq)/(Kgap_TAL1*Ks7p_TAL1)/((1+GAP/Kgap_TAL1+F6P/Kf6p_TAL1)*(1+S7P/Ks7p_TAL1+E4P/Ke4p_TAL1))+NQM1*kcat_NQM1*(GAP*S7P-F6P*E4P/Keq)/(Kgap_NQM1*Ks7p_NQM1)/((1+GAP/Kgap_NQM1+F6P/Kf6p_NQM1)*(1+S7P/Ks7p_NQM1+E4P/Ke4p_NQM1)))
Kigap=10.0 mM; Keq=0.069 mM; Kf16bp=0.4507 mM; Kgap=2.4 mM; kcat=4.139 per s; Kdhap=2.0 mMReaction: F16bP => DHAP + GAP; FBA1, FBA1, F16bP, DHAP, GAP, Rate Law: cell*FBA1*kcat*(F16bP/Kf16bp-DHAP*GAP/(Kf16bp*Keq))/(1+F16bP/Kf16bp+DHAP/Kdhap+GAP/Kgap+F16bP*GAP/(Kf16bp*Kigap)+DHAP*GAP/(Kdhap*Kgap))
Vmax=13.2552 mM per s; Kiutp=0.11 mM; Kg1p=0.32 mM; Kutp=0.11 mM; Kiudg=0.0035 mMReaction: G1P + UTP => UDG; UGP1, UTP, G1P, UDG, Rate Law: cell*Vmax*UTP*G1P/(Kutp*Kg1p)/(Kiutp/Kutp+UTP/Kutp+G1P/Kg1p+UTP*G1P/(Kutp*Kg1p)+Kiutp/Kutp*UDG/Kiudg+G1P*UDG/(Kg1p*Kiudg))
Knadp_GND2=0.094 mM; Kru5p_GND1=0.1 mM; kcat_GND2=27.3 per s; Knadph_GND2=0.055 mM; kcat_GND1=28.0 per s; Knadp_GND1=0.094 mM; Kru5p_GND2=0.1 mM; Knadph_GND1=0.055 mM; Kp6g_GND2=0.115 mM; Kp6g_GND1=0.062 mMReaction: P6G + NADP => Ru5P + NADPH; GND1, GND2, GND1, P6G, NADP, Ru5P, NADPH, GND2, Rate Law: cell*(GND1*kcat_GND1*P6G*NADP/(Kp6g_GND1*Knadp_GND1)/((1+P6G/Kp6g_GND1+Ru5P/Kru5p_GND1)*(1+NADP/Knadp_GND1+NADPH/Knadph_GND1))+GND2*kcat_GND2*P6G*NADP/((1+P6G/Kp6g_GND2+Ru5P/Kru5p_GND2)*(1+NADP/Knadp_GND2+NADPH/Knadph_GND2)))
Kglc=0.9 mM; Vmax=3.35 mM per s; Ki=0.91 dimensionlessReaction: GLCx => GLC; GLCx, GLC, Rate Law: cell*Vmax*(GLCx-GLC)/Kglc/(1+GLCx/Kglc+GLC/Kglc+Ki*GLCx/Kglc*GLC/Kglc)
Vmax=0.883333333333333 mM per s; Kg3p=3.5 mMReaction: G3P => GLY; HOR2, RHR2, G3P, Rate Law: cell*Vmax*G3P/Kg3p/(1+G3P/Kg3p)
Keq=0.45 dimensionless; k=0.75 per_mM_per_sReaction: ADP => ATP + AMP; ADP, AMP, ATP, Rate Law: cell*k*(ADP*ADP-AMP*ATP/Keq)
Kg6p=1.0257 mM; kcat=487.36 per s; Keq=0.29 dimensionless; Kf6p=0.307 mMReaction: G6P => F6P; PGI1, PGI1, G6P, F6P, Rate Law: cell*PGI1*kcat*(G6P/Kg6p-F6P/(Kg6p*Keq))/(1+G6P/Kg6p+F6P/Kf6p)
Kpep_ENO1=0.5 mM; kcat_ENO2=19.87 per s; Kpep_ENO2=0.5 mM; Kp2g_ENO2=0.104 mM; Kp2g_ENO1=0.043 mM; Keq=6.7 dimensionless; kcat_ENO1=7.6 per sReaction: P2G => PEP; ENO1, ENO2, ENO1, P2G, PEP, ENO2, Rate Law: cell*ENO1*kcat_ENO1*(P2G/Kp2g_ENO1-PEP/(Kp2g_ENO1*Keq))/(1+P2G/Kp2g_ENO1+PEP/Kpep_ENO1)+cell*ENO2*kcat_ENO2*(P2G/Kp2g_ENO2-PEP/(Kp2g_ENO2*Keq))/(1+P2G/Kp2g_ENO2+PEP/Kpep_ENO2)
kcat=20.146 per s; Kiatp=9.3 mM; Kpyr=21.0 mM; Katp=1.5 mM; Kf16p=0.2 mM; Kpep=0.281 mM; L0=100.0 dimensionless; Keq=6500.0 dimensionless; Kadp=0.243 mMReaction: ADP + PEP => ATP + PYR; CDC19, F16bP, CDC19, PEP, ADP, PYR, ATP, F16bP, Rate Law: cell*CDC19*kcat*(PEP*ADP-PYR*ATP/Keq)/(Kpep*Kadp)/((1+PEP/Kpep+PYR/Kpyr+L0*(ATP/Kiatp+1)/(F16bP/Kf16p+1))*(1+ADP/Kadp+ATP/Katp))
Ke4p_TAL = 0.946 mM; Keq=1.2 dimensionless; kcat=40.5 per s; Kgap_TAL = 0.1 mM; Ks7p_TAL = 0.15 mM; Kr5p_TAL = 0.235 mM; Kx5p_TAL = 0.67 mM; Kf6p_TAL = 1.1 mMReaction: X5P + R5P => GAP + S7P; TKL1, E4P, F6P, TKL1, X5P, R5P, GAP, S7P, E4P, F6P, Rate Law: cell*TKL1*kcat*(X5P*R5P-GAP*S7P/Keq)/(Kx5p_TAL*Kr5p_TAL)/((1+X5P/Kx5p_TAL+GAP/Kgap_TAL)*(1+E4P/Ke4p_TAL+F6P/Kf6p_TAL+R5P/Kr5p_TAL+S7P/Ks7p_TAL))
Kpyr_PDC1=8.5 mM; Kpyr_PDC6=2.92 mM; Kpyr_PDC5=7.08 mM; kcat_PDC6=9.21 per s; kcat_PDC5=10.32 per s; kcat_PDC1=12.14 per sReaction: PYR => AcAld; PDC1, PDC5, PDC6, PDC1, PYR, PDC5, PDC6, Rate Law: cell*PDC1*kcat_PDC1*PYR/Kpyr_PDC1/(1+PYR/Kpyr_PDC1)+cell*PDC5*kcat_PDC5*PYR/Kpyr_PDC5/(1+PYR/Kpyr_PDC5)+cell*PDC6*kcat_PDC6*PYR/Kpyr_PDC6/(1+PYR/Kpyr_PDC6)
Kinad=0.92 mM; Knadh=0.11 mM; kcat=176.0 per s; Kacald=0.4622 mM; Kinadh=0.031 mM; Keq=14492.7536231884 dimensionless; Kietoh=90.0 mM; Knad=0.17 mM; Kiacald=1.1 mM; Ketoh=17.0 mMReaction: AcAld + NADH => EtOH + NAD; ADH1, ADH1, AcAld, NADH, EtOH, NAD, Rate Law: cell*ADH1*kcat*(AcAld*NADH/(Kacald*Kinadh)-EtOH*NAD/(Kacald*Kinadh*Keq))/(1+NADH/Kinadh+AcAld*Knadh/(Kinadh*Kacald)+EtOH*Knad/(Kinad*Ketoh)+NAD/Kinad+AcAld*NADH/(Kinadh*Kacald)+NADH*EtOH*Knad/(Kinadh*Kinad*Ketoh)+AcAld*NAD*Knadh/(Kinadh*Kinad*Kacald)+EtOH*NAD/(Ketoh*Kinad)+AcAld*NADH*EtOH/(Kinadh*Kietoh*Kacald)+AcAld*EtOH*NAD/(Kiacald*Kinad*Ketoh))
Cf16=0.397 dimensionless; Kadp=1.0 mM; Katp=0.71 mM; Kiatp=0.65 mM; gR=5.12 dimensionless; Keq=800.0 dimensionless; Cf26=0.0174 dimensionless; Kf26=6.82E-4 mM; Kf6p=0.1 mM; Kf16=0.111 mM; kcat=209.6 per s; Camp=0.0845 dimensionless; L0=0.66 dimensionless; Ciatp=100.0 dimensionless; Kamp=0.0995 mM; Catp=3.0 dimensionlessReaction: ATP + F6P => ADP + F16bP; AMP, F26bP, PFK1, PFK2, PFK2, F6P, ATP, F16bP, ADP, AMP, F26bP, Rate Law: cell*PFK2*kcat*gR*F6P/Kf6p*ATP/Katp*(1-F16bP*ADP/(F6P*ATP)/Keq)*(1+F6P/Kf6p+ATP/Katp+gR*F6P/Kf6p*ATP/Katp+F16bP/Kf16+ADP/Kadp+gR*F16bP/Kf16*ADP/Kadp)/((1+F6P/Kf6p+ATP/Katp+gR*F6P/Kf6p*ATP/Katp+F16bP/Kf16+ADP/Kadp+gR*F16bP/Kf16*ADP/Kadp)^2+L0*((1+Ciatp*ATP/Kiatp)/(1+ATP/Kiatp))^2*((1+Camp*AMP/Kamp)/(1+AMP/Kamp))^2*((1+Cf26*F26bP/Kf26+Cf16*F16bP/Kf16)/(1+F26bP/Kf26+F16bP/Kf16))^2*(1+Catp*ATP/Katp)^2)
Keq=0.00533412710224736 dimensionless; Kgap_TDH3=0.423 mM; Knadh_TDH3=0.06 mM; Kbpg_TDH3=0.909 mM; kcat_TDH3=18.162 per s; Knad_TDH3=0.09 mM; Knad_TDH1=0.09 mM; Kgap_TDH1=0.495 mM; Kbpg_TDH1=0.0098 mM; Knadh_TDH1=0.06 mM; kcat_TDH1=19.12 per sReaction: GAP + NAD => BPG + NADH; TDH1, TDH3, TDH1, GAP, NAD, BPG, NADH, TDH3, Rate Law: cell*TDH1*kcat_TDH1*(GAP*NAD/(Kgap_TDH1*Knad_TDH1)-BPG*NADH/(Kgap_TDH1*Knad_TDH1*Keq))/((1+GAP/Kgap_TDH1+BPG/Kbpg_TDH1)*(1+NAD/Knad_TDH1+NADH/Knadh_TDH1))+cell*TDH3*kcat_TDH3*(GAP*NAD/(Kgap_TDH3*Knad_TDH3)-BPG*NADH/(Kgap_TDH3*Knad_TDH3*Keq))/((1+GAP/Kgap_TDH3+BPG/Kbpg_TDH3)*(1+NAD/Knad_TDH3+NADH/Knadh_TDH3))
kcat=47.1 per s; Keq=10.0 dimensionless; Ke4p_TAL = 0.946 mM; Kgap_TAL = 0.1 mM; Ks7p_TAL = 0.15 mM; Kr5p_TAL = 0.235 mM; Kx5p_TAL = 0.67 mM; Kf6p_TAL = 1.1 mMReaction: X5P + E4P => GAP + F6P; TKL1, R5P, S7P, TKL1, X5P, E4P, GAP, F6P, R5P, S7P, Rate Law: cell*TKL1*kcat*(X5P*E4P-GAP*F6P/Keq)/(Kx5p_TAL*Ke4p_TAL)/((1+X5P/Kx5p_TAL+GAP/Kgap_TAL)*(1+E4P/Ke4p_TAL+F6P/Kf6p_TAL+R5P/Kr5p_TAL+S7P/Ks7p_TAL))
Kt6p=0.5 mM; Vmax=2.33999999999999 mM per sReaction: T6P => TRH; TPS1, TPS2, T6P, Rate Law: cell*Vmax*T6P/Kt6p/(1+T6P/Kt6p)
Keq=0.045 dimensionless; Kigap=35.1 mM; Kdhap=6.454 mM; kcat=564.38 per s; Kgap=5.25 mMReaction: DHAP => GAP; TPI1, TPI1, DHAP, GAP, Rate Law: cell*TPI1*kcat/Kdhap*(DHAP-GAP/Keq)/(1+DHAP/Kdhap+GAP/Kgap*(1+(GAP/Kigap)^4))

States:

NameDescription
ACE[acetate]
ATP[ATP; ATP]
G1P[D-glucopyranose 1-phosphate]
F16bP[beta-D-fructofuranose 1,6-bisphosphate; beta-D-Fructose 1,6-bisphosphate]
GLC[D-glucopyranose; D-Glucose]
GLY[glycerol; Glycerol]
AMP[AMP; AMP]
DHAP[dihydroxyacetone phosphate; Glycerone phosphate]
NADPH[NADPH(4-); NADPH]
P2G[2-phospho-D-glyceric acid; 2-Phospho-D-glycerate]
T6P[alpha,alpha-trehalose 6-phosphate]
P3G[3-phospho-D-glyceric acid; 3-Phospho-D-glycerate]
P6G[6-phosphonatooxy-D-gluconate; 6-Phospho-D-gluconate]
UTP[UTP(4-)]
UDG[UDP-D-glucose]
GLCx[D-glucopyranose; D-Glucose]
NADH[NADH; NADH]
PYR[pyruvate; Pyruvate]
AcAld[acetaldehyde; Acetaldehyde]
R5P[alpha-D-ribofuranose 5-phosphate; alpha-D-Ribose 5-phosphate]
NADP[NADP(3-); NADP+]
EtOH[ethanol; Ethanol]
BPG[3-phospho-D-glyceroyl dihydrogen phosphate; 3-Phospho-D-glyceroyl phosphate]
X5P[D-xylulose 5-phosphate(2-); D-Xylulose 5-phosphate]
F6P[beta-D-fructofuranose 6-phosphate; beta-D-Fructose 6-phosphate]
S7P[sedoheptulose 7-phosphate(2-); Sedoheptulose 7-phosphate]
GAP[D-glyceraldehyde 3-phosphate; D-Glyceraldehyde 3-phosphate]
TRH[alpha,alpha-trehalose; alpha,alpha-Trehalose]
UDP[UDP]
E4P[D-erythrose 4-phosphate(2-); D-Erythrose 4-phosphate]
Ru5P[D-ribulose 5-phosphate(2-); D-Ribulose 5-phosphate]
G6P[alpha-D-glucose 6-phosphate; alpha-D-Glucose 6-phosphate]
PEP[phosphoenolpyruvate; Phosphoenolpyruvate]
NAD[NAD(+); NAD+]
ADP[ADP; ADP]
G6L[6-O-phosphonato-D-glucono-1,5-lactone(2-); D-Glucono-1,5-lactone 6-phosphate]
G3P[sn-glycerol 3-phosphate; sn-Glycerol 3-phosphate]

Messiha2013 - Pentose phosphate pathway model: BIOMD0000000502v0.0.1

Messiha2013 - Pentose phosphate pathway modelThis model describes the dynamic behaviour of the pentose phosphate pathway…

Details

We present the quantification and kinetic characterisation of the enzymes of the pentose phosphate pathway in Saccharomyces cerevisiae. The data are combined into a mathematical model that describes the dynamics of this system and allows for the predicting changes in metabolite concentrations and fluxes in response to perturbations. We use the model to study the response of yeast to a glucose pulse. We then combine the model with an existing glycolysis one to study the effect of oxidative stress on carbohydrate metabolism. The combination of these two models was made possible by the standardized enzyme kinetic experiments carried out in both studies. This work demonstrates the feasibility of constructing larger network models by merging smaller pathway models. link: http://identifiers.org/doi/10.7287/peerj.preprints.146v2

Parameters:

NameDescription
k=1.0 per sReaction: R5P => ; R5P, Rate Law: cell*k*R5P
Knadp=0.045 mM; kcat=189.0 per s; Kg6p=0.042 mM; Knadph=0.017 mM; Kg6l=0.01 mMReaction: G6P + NADP => G6L + NADPH; ZWF1, ZWF1, G6P, NADP, G6L, NADPH, Rate Law: cell*ZWF1*kcat*G6P*NADP/(Kg6p*Knadp)/((1+G6P/Kg6p+G6L/Kg6l)*(1+NADP/Knadp+NADPH/Knadph))
Kru5p=2.47 mM; kcat=335.0 per s; Kiru5p=9.88 mM; Keq=4.0 dimensionless; Kr5p=5.7 mMReaction: Ru5P => R5P; RKI1, RKI1, Ru5P, R5P, Rate Law: cell*RKI1*kcat*(Ru5P-R5P/Keq)/Kru5p/(1+Ru5P/Kru5p+R5P/Kr5p)
kcat=47.1 per s; Keq=10.0 dimensionless; Ke4p_TAL = 0.946 mM; Kgap_TAL = 0.1 mM; Ks7p_TAL = 0.15 mM; Kr5p_TAL = 0.235 mM; Kx5p_TAL = 0.67 mM; Kf6p_TAL = 1.1 mMReaction: X5P + E4P => GAP + F6P; TKL1, R5P, S7P, TKL1, X5P, E4P, GAP, F6P, R5P, S7P, Rate Law: cell*TKL1*kcat*(X5P*E4P-GAP*F6P/Keq)/(Kx5p_TAL*Ke4p_TAL)/((1+X5P/Kx5p_TAL+GAP/Kgap_TAL)*(1+E4P/Ke4p_TAL+F6P/Kf6p_TAL+R5P/Kr5p_TAL+S7P/Ks7p_TAL))
Kg6l=0.8 mM; kcat=4.3 per s; Kp6g=0.5 mMReaction: G6L => P6G; SOL3, SOL3, G6L, P6G, Rate Law: cell*SOL3*kcat*G6L/Kg6l/(1+G6L/Kg6l+P6G/Kp6g)
sum_NADP = 0.33 mMReaction: NADP = sum_NADP-NADPH, Rate Law: missing
Kgap_TAL1=0.272 mM; Ke4p_NQM1=0.305 mM; kcat_NQM1=0.694 per s; Kf6p_TAL1=1.44 mM; kcat_TAL1=0.694 per s; Kgap_NQM1=0.272 mM; Ks7p_NQM1=0.786 mM; Kf6p_NQM1=1.04 mM; Ke4p_TAL1=0.362 mM; Keq=1.05 dimensionless; Ks7p_TAL1=0.786 mMReaction: GAP + S7P => F6P + E4P; TAL1, NQM1, TAL1, GAP, S7P, F6P, E4P, NQM1, Rate Law: cell*(TAL1*kcat_TAL1*(GAP*S7P-F6P*E4P/Keq)/(Kgap_TAL1*Ks7p_TAL1)/((1+GAP/Kgap_TAL1+F6P/Kf6p_TAL1)*(1+S7P/Ks7p_TAL1+E4P/Ke4p_TAL1))+NQM1*kcat_NQM1*(GAP*S7P-F6P*E4P/Keq)/(Kgap_NQM1*Ks7p_NQM1)/((1+GAP/Kgap_NQM1+F6P/Kf6p_NQM1)*(1+S7P/Ks7p_NQM1+E4P/Ke4p_NQM1)))
Ke4p_TAL = 0.946 mM; Keq=1.2 dimensionless; kcat=40.5 per s; Kgap_TAL = 0.1 mM; Ks7p_TAL = 0.15 mM; Kr5p_TAL = 0.235 mM; Kx5p_TAL = 0.67 mM; Kf6p_TAL = 1.1 mMReaction: X5P + R5P => GAP + S7P; TKL1, E4P, F6P, TKL1, X5P, R5P, GAP, S7P, E4P, F6P, Rate Law: cell*TKL1*kcat*(X5P*R5P-GAP*S7P/Keq)/(Kx5p_TAL*Kr5p_TAL)/((1+X5P/Kx5p_TAL+GAP/Kgap_TAL)*(1+E4P/Ke4p_TAL+F6P/Kf6p_TAL+R5P/Kr5p_TAL+S7P/Ks7p_TAL))
Kx5p=7.7 mM; Keq=1.4 dimensionless; kcat=4020.0 per s; Kru5p=5.97 mMReaction: Ru5P => X5P; RPE1, RPE1, Ru5P, X5P, Rate Law: cell*RPE1*kcat*(Ru5P-X5P/Keq)/Kru5p/(1+Ru5P/Kru5p+X5P/Kx5p)
Knadp_GND2=0.094 mM; Kru5p_GND1=0.1 mM; kcat_GND2=27.3 per s; Knadph_GND2=0.055 mM; kcat_GND1=28.0 per s; Knadp_GND1=0.094 mM; Kru5p_GND2=0.1 mM; Knadph_GND1=0.055 mM; Kp6g_GND2=0.115 mM; Kp6g_GND1=0.062 mMReaction: P6G + NADP => Ru5P + NADPH; GND1, GND2, GND1, P6G, NADP, Ru5P, NADPH, GND2, Rate Law: cell*(GND1*kcat_GND1*P6G*NADP/(Kp6g_GND1*Knadp_GND1)/((1+P6G/Kp6g_GND1+Ru5P/Kru5p_GND1)*(1+NADP/Knadp_GND1+NADPH/Knadph_GND1))+GND2*kcat_GND2*P6G*NADP/((1+P6G/Kp6g_GND2+Ru5P/Kru5p_GND2)*(1+NADP/Knadp_GND2+NADPH/Knadph_GND2)))

States:

NameDescription
NADP[NADP(3-); NADP+]
R5P[alpha-D-ribofuranose 5-phosphate; alpha-D-Ribose 5-phosphate]
X5P[D-xylulose 5-phosphate(2-); D-Xylulose 5-phosphate]
F6P[D-fructose 6-phosphate(2-); D-Fructose 6-phosphate]
S7P[sedoheptulose 7-phosphate(2-); Sedoheptulose 7-phosphate]
E4P[D-erythrose 4-phosphate(2-); D-Erythrose 4-phosphate]
GAP[glyceraldehyde 3-phosphate(2-); Glyceraldehyde 3-phosphate]
G6P[D-glucose 6-phosphate; D-Glucose 6-phosphate]
NADPH[NADPH(4-); NADPH]
Ru5P[D-ribulose 5-phosphate(2-); D-Ribulose 5-phosphate]
P6G[6-phosphonatooxy-D-gluconate; 6-Phospho-D-gluconate]
G6L[6-O-phosphonato-D-glucono-1,5-lactone(2-); D-Glucono-1,5-lactone 6-phosphate]

Meyer1991_CalciumSpike_ICC: BIOMD0000000224v0.0.1

This a model from the article: Calcium spiking. Meyer T, Stryer L Annu Rev Biophys Biophys Chem1991:20:153-74…

Details

link: http://identifiers.org/pubmed/1867714

Parameters:

NameDescription
E = 1.0 microM4Reaction: => g; CaI, Rate Law: E*(CaI*0.01)^4*(1-g)
A = 20.0 psec; L = 0.01 psec; k1 = 0.5 microMReaction: CaS => CaI; IP3, g, Rate Law: (1-g)*(A*(IP3*0.5)^4/(IP3*0.5+k1)^4+L)*CaS
F = 0.02 psecReaction: g => ; CaI, Rate Law: F
k2 = 0.15 microM; B = 40.0 microMpsecReaction: CaI => CaS, Rate Law: B*(CaI*0.01)^2/((CaI*0.01)^2+k2^2)
D = 2.0 psecReaction: IP3 =>, Rate Law: D*IP3*0.5
k3 = 1.0 microM; R = 0.09; C = 1.1 microMpsecReaction: => IP3; CaI, Rate Law: C*(1-k3/(CaI*0.01+k3)*1/(1+R))

States:

NameDescription
IP3[1D-myo-inositol 1,4,5-trisphosphate; N-(6-Aminohexanoyl)-6-aminohexanoate]
g[calcium channel inhibitor activity]
CaI[calcium(2+); Calcium cation]
CaS[calcium(2+); Calcium cation]

Miao2010 - Innate and adaptive immune responses to primary Influenza A Virus infection: BIOMD0000000546v0.0.1

Miao2010 - Innate and adaptive immune responses to primary Influenza A Virus infectionThis model is described in the art…

Details

Seasonal and pandemic influenza A virus (IAV) continues to be a public health threat. However, we lack a detailed and quantitative understanding of the immune response kinetics to IAV infection and which biological parameters most strongly influence infection outcomes. To address these issues, we use modeling approaches combined with experimental data to quantitatively investigate the innate and adaptive immune responses to primary IAV infection. Mathematical models were developed to describe the dynamic interactions between target (epithelial) cells, influenza virus, cytotoxic T lymphocytes (CTLs), and virus-specific IgG and IgM. IAV and immune kinetic parameters were estimated by fitting models to a large data set obtained from primary H3N2 IAV infection of 340 mice. Prior to a detectable virus-specific immune response (before day 5), the estimated half-life of infected epithelial cells is approximately 1.2 days, and the half-life of free infectious IAV is approximately 4 h. During the adaptive immune response (after day 5), the average half-life of infected epithelial cells is approximately 0.5 days, and the average half-life of free infectious virus is approximately 1.8 min. During the adaptive phase, model fitting confirms that CD8(+) CTLs are crucial for limiting infected cells, while virus-specific IgM regulates free IAV levels. This may imply that CD4 T cells and class-switched IgG antibodies are more relevant for generating IAV-specific memory and preventing future infection via a more rapid secondary immune response. Also, simulation studies were performed to understand the relative contributions of biological parameters to IAV clearance. This study provides a basis to better understand and predict influenza virus immunity. link: http://identifiers.org/pubmed/20410284

Parameters:

NameDescription
rho_E = 6.2E-8 substanceReaction: s4 => s1; s1, s1, Rate Law: rho_E*s1
pi_a = 100.0 substanceReaction: s7 => s3; s2, s2, s2, Rate Law: pi_a*s2
beta_a = 2.4E-6 substanceReaction: s1 => s2; s3, s1, s3, s1, s3, Rate Law: beta_a*s1*s3
c_V = 4.2 substanceReaction: s3 => s6; s3, s3, Rate Law: c_V*s3
delta_Es = 0.6 substanceReaction: s2 => s5; s2, s2, Rate Law: delta_Es*s2

States:

NameDescription
s1[Epithelial cell]
s5s5
s6s6
s7s7
s2[Epithelial cell]
s4s4
s3[Influenza A virus (strain A/X-31 H3N2)]

Miao2014 - Dynamics and migratory pathways of virus-specific antibody-secreting cell populations: MODEL1411130000v0.0.1

</head> <body>This model is from the paper available at http://www.ncbi.nlm.nih.gov/pubmed/25171166. The original mode…

Details

The B cell response to influenza infection of the respiratory tract contributes to viral clearance and establishes profound resistance to reinfection by related viruses. Numerous studies have measured virus-specific antibody-secreting cell (ASC) frequencies in different anatomical compartments after influenza infection and provided a general picture of the kinetics of ASC formation and dispersion. However, the dynamics of ASC populations are difficult to determine experimentally and have received little attention. Here, we applied mathematical modeling to investigate the dynamics of ASC growth, death, and migration over the 2-week period following primary influenza infection in mice. Experimental data for model fitting came from high frequency measurements of virus-specific IgM, IgG, and IgA ASCs in the mediastinal lymph node (MLN), spleen, and lung. Model construction was based on a set of assumptions about ASC gain and loss from the sampled sites, and also on the directionality of ASC trafficking pathways. Most notably, modeling results suggest that differences in ASC fate and trafficking patterns reflect the site of formation and the expressed antibody class. Essentially all early IgA ASCs in the MLN migrated to spleen or lung, whereas cell death was likely the major reason for IgM and IgG ASC loss from the MLN. In contrast, the spleen contributed most of the IgM and IgG ASCs that migrated to the lung, but essentially none of the IgA ASCs. This finding points to a critical role for regional lymph nodes such as the MLN in the rapid generation of IgA ASCs that seed the lung. Results for the MLN also suggest that ASC death is a significant early feature of the B cell response. Overall, our analysis is consistent with accepted concepts in many regards, but it also indicates novel features of the B cell response to influenza that warrant further investigation. link: http://identifiers.org/pubmed/25171166

Middleton2012_GibberellinSignalling: BIOMD0000000422v0.0.1

This model is from the article: Mathematical modeling elucidates the role of transcriptional feedback in gibberellin s…

Details

The hormone gibberellin (GA) is a key regulator of plant growth. Many of the components of the gibberellin signal transduction [e.g., GIBBERELLIN INSENSITIVE DWARF 1 (GID1) and DELLA], biosynthesis [e.g., GA 20-oxidase (GA20ox) and GA3ox], and deactivation pathways have been identified. Gibberellin binds its receptor, GID1, to form a complex that mediates the degradation of DELLA proteins. In this way, gibberellin relieves DELLA-dependent growth repression. However, gibberellin regulates expression of GID1, GA20ox, and GA3ox, and there is also evidence that it regulates DELLA expression. In this paper, we use integrated mathematical modeling and experiments to understand how these feedback loops interact to control gibberellin signaling. Model simulations are in good agreement with in vitro data on the signal transduction and biosynthesis pathways and in vivo data on the expression levels of gibberellin-responsive genes. We find that GA-GID1 interactions are characterized by two timescales (because of a lid on GID1 that can open and close slowly relative to GA-GID1 binding and dissociation). Furthermore, the model accurately predicts the response to exogenous gibberellin after a number of chemical and genetic perturbations. Finally, we investigate the role of the various feedback loops in gibberellin signaling. We find that regulation of GA20ox transcription plays a significant role in both modulating the level of endogenous gibberellin and generating overshoots after the removal of exogenous gibberellin. Moreover, although the contribution of other individual feedback loops seems relatively small, GID1 and DELLA transcriptional regulation acts synergistically with GA20ox feedback. link: http://identifiers.org/pubmed/22523240

Parameters:

NameDescription
gammaGA20ox=3.514 substanceReaction: s27 => s6; s27, Rate Law: gammaGA20ox*s27
ua1=10.0 substanceReaction: s62 + s16 => s45; s62, s16, Rate Law: ua1*s62*s16
kd15=0.008827838388125 substanceReaction: s31 => s24 + s27; s31, Rate Law: kd15*s31
thetaGA20ox=0.6383 substance; muGA20ox = 0.047770755070625 substanceReaction: s11 => s39; s16, s16, Rate Law: muGA20ox*s16/(s16+thetaGA20ox)
deltaGA20ox=0.192990314378105 substanceReaction: s6 => s27; s39, s39, Rate Law: deltaGA20ox*s39
um=6.92208879449283 substanceReaction: s45 => s62 + s22; s45, Rate Law: um*s45
muGID = 0.045708818961487 substanceReaction: s42 => s15; s42, Rate Law: muGID*s42
muGA3ox = 0.102600014140148 substanceReaction: s41 => s35; s41, Rate Law: muGA3ox*s41
ud2=2.8184 substanceReaction: s36 => s62 + s16; s36, Rate Law: ud2*s36
muDELLA = 0.070794578438414 substanceReaction: s40 => s34; s40, Rate Law: muDELLA*s40
muGA = 0.290804218727464 substanceReaction: s24 => s68; s24, Rate Law: muGA*s24
thetaGA3ox=0.0082 substance; muGA3ox = 0.102600014140148 substanceReaction: s35 => s41; s16, s16, Rate Law: muGA3ox*s16/(s16+thetaGA3ox)
kd9=0.008827838388125 substanceReaction: s29 => s26 + s28; s29, Rate Law: kd9*s29
ka24=3099.18915892587 substanceReaction: s25 + s27 => s30; s25, s27, Rate Law: ka24*s25*s27
la=1.35 substanceReaction: s1 + s2 => s65; s1, s2, Rate Law: la*s1*s2
ua2=316.2278 substanceReaction: s62 + s16 => s36; s62, s16, Rate Law: ua2*s62*s16
ka15=2073.22402517968 substanceReaction: s24 + s27 => s31; s24, s27, Rate Law: ka15*s24*s27
ud1=0.133045441797809 substanceReaction: s45 => s62 + s16; s45, Rate Law: ud1*s45
kd12=2.67298621993027 substanceReaction: s32 => s23 + s27; s32, Rate Law: kd12*s32
km9=763.777072066507 substanceReaction: s29 => s28 + s1; s29, Rate Law: km9*s29
deltaDELLA=5.27749140286577E-4 substanceReaction: s7 => s16; s40, s40, Rate Law: deltaDELLA*s40
q=0.025118864315096 substanceReaction: s65 => s62; s65, Rate Law: q*s65
deltaGA3ox=0.019299031437811 substanceReaction: s5 => s28; s41, s41, Rate Law: deltaGA3ox*s41
km24=2.58846077319221 substanceReaction: s30 => s27 + s26; s30, Rate Law: km24*s30
thetaGID=5.5995E-4 substance; muGID = 0.045708818961487 substanceReaction: s15 => s42; s16, s16, Rate Law: muGID*s16/(s16+thetaGID)
km12=198.80427707769 substanceReaction: s32 => s27 + s24; s32, Rate Law: km12*s32
p=0.077624711662869 substanceReaction: s62 => s65; s62, Rate Law: p*s62
deltaGID=19.2990314378105 substanceReaction: s33 => s2; s42, s42, Rate Law: deltaGID*s42
omegaGA4 = 0.0 substance; A1=0.0307 substance; Pmem = 2.66664 substanceReaction: s66 => s1, Rate Law: Pmem*A1*omegaGA4
gammaGID=3.514 substanceReaction: s2 => s33; s2, Rate Law: gammaGID*s2
thetaDELLA=0.01 substance; muDELLA = 0.070794578438414 substanceReaction: s34 => s40; s16, s16, Rate Law: muDELLA*thetaDELLA/(s16+thetaDELLA)
ka12=2904.11853677638 substanceReaction: s23 + s27 => s32; s23, s27, Rate Law: ka12*s23*s27
muGA20ox = 0.047770755070625 substanceReaction: s39 => s11; s39, Rate Law: muGA20ox*s39
kd24=0.01588492846351 substanceReaction: s30 => s25 + s27; s30, Rate Law: kd24*s30
omegaGA12 = 0.006602803853512 substanceReaction: s3 => s23, Rate Law: omegaGA12
km15=763.777072066507 substanceReaction: s31 => s27 + s25; s31, Rate Law: km15*s31
Pmem = 2.66664 substance; B1=3.9795E-4 substance; muGA = 0.290804218727464 substanceReaction: s1 => s71; s1, Rate Law: (muGA+Pmem*B1)*s1
ld=2.84315148627376 substanceReaction: s65 => s1 + s2; s65, Rate Law: ld*s65
ka9=2073.22402517968 substanceReaction: s26 + s28 => s29; s26, s28, Rate Law: ka9*s26*s28
gammaGA3ox=3.514 substanceReaction: s28 => s5; s28, Rate Law: gammaGA3ox*s28

States:

NameDescription
s23[gibberellin A12]
s5GA3ox_source
s24[gibberellin A15 (diacid form)]
s40[DELLA protein GAI]
s35ga3ox_source
s7DELLA_source
s71sa1_degraded
s31[gibberellin A15 (diacid form); GA20OX1]
s34della_source
s36[gibberellin A4; Gibberellin receptor GID1A; DELLA protein GAI]
s6GA20ox_source
s32[gibberellin A12; GA20OX1]
s22[DELLA protein GAI]
s70sa8_degraded
s11ga20ox_source
s15gid_source
s45[gibberellin A4; DELLA protein GAI; Gibberellin receptor GID1A]
s69sa7_degraded
s3GA12_source
s1[gibberellin A4]
s67sa5_degraded
s41[GA3OX3]
s68sa6_degraded
s25[gibberellin A24]
s2[Gibberellin receptor GID1A]
s33GID_source
s16[DELLA protein GAI]
s30[gibberellin A24; GA20OX1]
s26[gibberellin A9]
s62[gibberellin A4; Gibberellin receptor GID1A]
s42[Gibberellin receptor GID1A]
s28[GA3OX3]
s39[GA20OX1]
s65[gibberellin A4; Gibberellin receptor GID1A]
s66GA4_source
s29[gibberellin A9; GA3OX3]
s27[GA20OX1]

Millard2016 - E. coli central carbon and energy metabolism: MODEL1505110000v0.0.1

Millard2016 - E. coli central carbon and energy metabolismThis model is described in the article: [Metabolic regulation…

Details

The metabolism of microorganisms is regulated through two main mechanisms: changes of enzyme capacities as a consequence of gene expression modulation (“hierarchical control”) and changes of enzyme activities through metabolite-enzyme interactions. An increasing body of evidence indicates that hierarchical control is insufficient to explain metabolic behaviors, but the system-wide impact of metabolic regulation remains largely uncharacterized. To clarify its role, we developed and validated a detailed kinetic model of Escherichia coli central metabolism that links growth to environment. Metabolic control analyses confirm that the control is widely distributed across the network and highlight strong interconnections between all the pathways. Exploration of the model solution space reveals that several robust properties emerge from metabolic regulation, from the molecular level (e.g. homeostasis of total metabolite pool) to the overall cellular physiology (e.g. coordination of carbon uptake, catabolism, energy and redox production, and growth), while allowing a large degree of flexibility at most individual metabolic steps. These properties have important physiological implications for E. coli and significantly expand the self-regulating capacities of its metabolism. link: http://identifiers.org/doi/10.1371/journal.pcbi.1005396

Millard2020 - Kinetic model of Glucose and Acetate metabolisms in E. coli: MODEL2005050001v0.0.1

Calibrated kinetic model of glucose and acetate metabolisms of Escherichia coli, as detailed in Millard et al., 2020 (DO…

Details

Overflow metabolism refers to the production of seemingly wasteful by-products by cells during growth on glucose even when oxygen is abundant. Two theories have been proposed to explain acetate overflow in Escherichia coli - global control of the central metabolism and local control of the acetate pathway - but neither accounts for all observations. Here, we develop a kinetic model of E. coli metabolism that quantitatively accounts for observed behaviors and successfully predicts the response of E. coli to new perturbations. We reconcile these theories and clarify the origin, control and regulation of the acetate flux. We also find that, in turns, acetate regulates glucose metabolism by coordinating the expression of glycolytic and TCA genes. Acetate should not be considered a wasteful end-product since it is also a co-substrate and a global regulator of glucose metabolism in E. coli. This has broad implications for our understanding of overflow metabolism. link: http://identifiers.org/doi/10.1101/2020.08.18.255356

Milne2011 - Genome-scale metabolic network of Clostridium beijerinckii (iCB925): MODEL1507180034v0.0.1

Milne2011 - Genome-scale metabolic network of Clostridium beijerinckii (iCB925)This model is described in the article:…

Details

BACKGROUND: Solventogenic clostridia offer a sustainable alternative to petroleum-based production of butanol–an important chemical feedstock and potential fuel additive or replacement. C. beijerinckii is an attractive microorganism for strain design to improve butanol production because it (i) naturally produces the highest recorded butanol concentrations as a byproduct of fermentation; and (ii) can co-ferment pentose and hexose sugars (the primary products from lignocellulosic hydrolysis). Interrogating C. beijerinckii metabolism from a systems viewpoint using constraint-based modeling allows for simulation of the global effect of genetic modifications. RESULTS: We present the first genome-scale metabolic model (iCM925) for C. beijerinckii, containing 925 genes, 938 reactions, and 881 metabolites. To build the model we employed a semi-automated procedure that integrated genome annotation information from KEGG, BioCyc, and The SEED, and utilized computational algorithms with manual curation to improve model completeness. Interestingly, we found only a 34% overlap in reactions collected from the three databases–highlighting the importance of evaluating the predictive accuracy of the resulting genome-scale model. To validate iCM925, we conducted fermentation experiments using the NCIMB 8052 strain, and evaluated the ability of the model to simulate measured substrate uptake and product production rates. Experimentally observed fermentation profiles were found to lie within the solution space of the model; however, under an optimal growth objective, additional constraints were needed to reproduce the observed profiles–suggesting the existence of selective pressures other than optimal growth. Notably, a significantly enriched fraction of actively utilized reactions in simulations–constrained to reflect experimental rates–originated from the set of reactions that overlapped between all three databases (P = 3.52 × 10-9, Fisher's exact test). Inhibition of the hydrogenase reaction was found to have a strong effect on butanol formation–as experimentally observed. CONCLUSIONS: Microbial production of butanol by C. beijerinckii offers a promising, sustainable, method for generation of this important chemical and potential biofuel. iCM925 is a predictive model that can accurately reproduce physiological behavior and provide insight into the underlying mechanisms of microbial butanol production. As such, the model will be instrumental in efforts to better understand, and metabolically engineer, this microorganism for improved butanol production. link: http://identifiers.org/pubmed/21846360

Mitchell2013 - Liver Iron Metabolism: BIOMD0000000498v0.0.1

Mitchell2013 - Liver Iron MetabolismThe model includes the core regulatory components of human liver iron metabolism. T…

Details

Iron is essential for all known life due to its redox properties; however, these same properties can also lead to its toxicity in overload through the production of reactive oxygen species. Robust systemic and cellular control are required to maintain safe levels of iron, and the liver seems to be where this regulation is mainly located. Iron misregulation is implicated in many diseases, and as our understanding of iron metabolism improves, the list of iron-related disorders grows. Recent developments have resulted in greater knowledge of the fate of iron in the body and have led to a detailed map of its metabolism; however, a quantitative understanding at the systems level of how its components interact to produce tight regulation remains elusive. A mechanistic computational model of human liver iron metabolism, which includes the core regulatory components, is presented here. It was constructed based on known mechanisms of regulation and on their kinetic properties, obtained from several publications. The model was then quantitatively validated by comparing its results with previously published physiological data, and it is able to reproduce multiple experimental findings. A time course simulation following an oral dose of iron was compared to a clinical time course study and the simulation was found to recreate the dynamics and time scale of the systems response to iron challenge. A disease state simulation of haemochromatosis was created by altering a single reaction parameter that mimics a human haemochromatosis gene (HFE) mutation. The simulation provides a quantitative understanding of the liver iron overload that arises in this disease. This model supports and supplements understanding of the role of the liver as an iron sensor and provides a framework for further modelling, including simulations to identify valuable drug targets and design of experiments to improve further our knowledge of this system. link: http://identifiers.org/pubmed/24244122

Parameters:

NameDescription
k1=3.209E-5Reaction: species_1 => ; species_1, species_1, Rate Law: compartment_1*k1*species_1
k1=22922.0Reaction: species_24 => species_2 + species_25; species_24, species_24, Rate Law: compartment_1*k1*species_24
K=3.0E-6; a=2.0; n=1.0Reaction: species_2 => species_43; species_4, species_4, species_2, species_4, species_2, Rate Law: a*species_4^n/(K^n+species_4^n)*species_2
k1=3.9438E11Reaction: species_8 + species_10 => species_18; species_8, species_10, species_8, species_10, Rate Law: compartment_3*k1*species_8^2*species_10
k1=108000.0Reaction: species_24 => species_26 + species_25; species_24, species_24, Rate Law: compartment_1*k1*species_24
k1=1.597E-5Reaction: species_6 => ; species_6, species_6, Rate Law: compartment_1*k1*species_6
k1=6.418E-5Reaction: species_8 => ; species_8, species_8, Rate Law: compartment_3*k1*species_8
n=5.0; a=2.315E-4; K=5.0E-9Reaction: species_4 => ; species_7, species_7, species_4, species_7, species_4, Rate Law: compartment_1*a*species_7^n/(K^n+species_7^n)*species_4
k1=4.0E-4Reaction: species_2 => ; species_2, species_2, Rate Law: compartment_1*k1*species_2
k1=0.024Reaction: species_19 => species_15 + species_43; species_19, species_19, Rate Law: compartment_3*k1*species_19
k1=4.71E10Reaction: species_2 + species_25 => species_24; species_2, species_25, species_2, species_25, Rate Law: compartment_1*k1*species_2*species_25
k1=69600.0Reaction: species_15 + species_43 => species_19; species_15, species_43, species_15, species_43, Rate Law: compartment_3*k1*species_15*species_43
k1=0.003535Reaction: species_16 => species_12 + species_43; species_16, species_16, Rate Law: compartment_3*k1*species_16
K=2.5E-6; a=3.2E-5; n=1.0Reaction: species_10 => ; species_43, species_43, species_10, species_43, species_10, Rate Law: compartment_3*a*(1-species_43^n/(K^n+species_43^n))*species_10
k1=5.6E-4Reaction: species_7 => ; species_7, species_7, Rate Law: compartment_1*k1*species_7
k1=1102000.0Reaction: species_9 + species_8 => species_17; species_9, species_8, species_9, species_8, Rate Law: compartment_3*k1*species_9*species_8
k1=8.37E-5Reaction: species_18 => ; species_18, species_18, Rate Law: compartment_3*k1*species_18
Km=1.78E-5; V=2.18E-5Reaction: species_5 => species_11; species_5, species_5, Rate Law: V*species_5/(Km+species_5)
k1=8.37E-7Reaction: species_17 => ; species_17, species_17, Rate Law: compartment_3*k1*species_17
k1=0.0018Reaction: species_18 => species_8 + species_10; species_18, species_18, Rate Law: compartment_3*k1*species_18
K=2.0E-6; C=17777.7Reaction: species_5 => species_2; species_1, species_1, species_5, species_1, species_5, Rate Law: compartment_1*species_1*C*species_5/(K+species_5)
a=1.0E-9; K=5.0E-6; n=1.0Reaction: => species_4; species_6, species_6, species_6, Rate Law: compartment_1*a*(1-species_6^n/(K^n+species_6^n))
k1=0.8333Reaction: species_16 => species_2 + species_3; species_16, species_16, Rate Law: k1*species_16
k1=837400.0Reaction: species_43 + species_3 => species_12; species_43, species_3, species_43, species_3, Rate Law: compartment_3*k1*species_43*species_3
k1=1.203E-5Reaction: species_25 => ; species_25, species_25, Rate Law: compartment_1*k1*species_25
K=1.0E-9; a=2.1432E-15Reaction: => species_1; species_5, species_5, species_5, Rate Law: compartment_1*a*species_5/(K+species_5)
k1=0.0061Reaction: species_15 => species_43 + species_10; species_15, species_15, Rate Law: compartment_3*k1*species_15
v=3.0E-11Reaction: => species_10, Rate Law: compartment_3*v
a=6.0E-12; K=1.0E-6; n=1.0Reaction: => species_3; species_6, species_6, species_6, Rate Law: compartment_3*a*species_6^n/(K^n+species_6^n)
v=2.3469E-11Reaction: => species_8, Rate Law: compartment_3*v
K=1.203E-5Reaction: species_26 => species_2; species_26, species_25, species_26, species_25, species_26, species_25, Rate Law: compartment_1*K*species_26/species_25*species_25
basal=0.0; a1=5.0E-12; n=5.0; K=1.35E-7; a=5.0E-12; K1=6.0E-7Reaction: => species_7; species_18, species_19, species_18, species_19, species_18, species_19, Rate Law: compartment_1*(basal+a*species_18^n/(K^n+species_18^n)+a1*species_19/(K1+species_19))
kloss=13.112Reaction: species_26 => species_2; species_26, species_25, species_26, species_25, species_26, species_25, Rate Law: compartment_1*species_26*kloss*(1+0.048*species_26/species_25/(1+species_26/species_25))
k1=0.08Reaction: species_17 => species_9 + species_8; species_17, species_17, Rate Law: compartment_3*k1*species_17
K=1.0E-6; n=1.0; a=4.0E-11Reaction: => species_6; species_2, species_2, species_2, Rate Law: compartment_1*a*(1-species_2^n/(K^n+species_2^n))
V=1.034E-5; Km=1.25E-4Reaction: species_11 => species_5; species_11, species_11, Rate Law: V*species_11/(Km+species_11)
a=2.312E-13; K=1.0E-6; n=1.0Reaction: => species_25; species_6, species_6, species_6, Rate Law: compartment_1*a*(1-species_6^n/(K^n+species_6^n))
k1=222390.0Reaction: species_43 + species_10 => species_15; species_43, species_10, species_43, species_10, Rate Law: compartment_3*k1*species_43*species_10
k1=121400.0Reaction: species_12 + species_43 => species_16; species_12, species_43, species_12, species_43, Rate Law: compartment_3*k1*species_12*species_43
k1=8.37E-6Reaction: species_3 => ; species_3, species_3, Rate Law: compartment_3*k1*species_3
k1=9.142E-4Reaction: species_12 => species_43 + species_3; species_12, species_12, Rate Law: compartment_3*k1*species_12

States:

NameDescription
species 9[Transferrin receptor protein 1; Hereditary hemochromatosis protein]
species 2[Iron; iron atom]
species 6[Cytoplasmic aconitate hydratase]
species 19[Transferrin receptor protein 2; Serotransferrin; Iron; iron atom]
species 10[Transferrin receptor protein 2]
species 11[Heme; ferroheme b]
species 1[Heme oxygenase 1]
species 18[Hereditary hemochromatosis protein; Transferrin receptor protein 2]
species 4[Solute carrier family 40 member 1]
species 16[Serotransferrin; Transferrin receptor protein 1; Iron; iron atom]
species 24[Ferritin light chain; Iron; iron atom]
species 43[Serotransferrin; Iron; iron atom]
species 3[Transferrin receptor protein 1]
species 25[Ferritin light chain]
species 8[Hereditary hemochromatosis protein]
species 17[Transferrin receptor protein 1; Hereditary hemochromatosis protein]
species 12[Serotransferrin; Transferrin receptor protein 2; Iron; iron atom]
species 7[Hepcidin]
species 5[Heme; ferroheme b]
species 15[Serotransferrin; Transferrin receptor protein 2; Iron; iron atom]
species 26[Iron; iron atom]

Mitrophanov2011 - Mode of action of rFVIIa on thrombin generation (Update to Hockin2002 Blood Coagulation Model): MODEL1806040001v0.0.1

Mathematical model of blood coagulation investigating effects of varied factor VIIa on thrombin generation. Model derive…

Details

INTRODUCTION: The therapeutic potential of a hemostatic agent can be assessed by investigating its effects on the quantitative parameters of thrombin generation. For recombinant activated factor VII (rFVIIa)–a promising hemostasis-inducing biologic–experimental studies addressing its effects on thrombin generation yielded disparate results. To elucidate the inherent ability of rFVIIa to modulate thrombin production, it is necessary to identify rFVIIa-induced effects that are compatible with the available biochemical knowledge about thrombin generation mechanisms. MATERIALS AND METHODS: The existing body of knowledge about coagulation biochemistry can be rigorously represented by a computational model that incorporates the known reactions and parameter values constituting the biochemical network. We used a thoroughly validated numerical model to generate activated factor VII (FVIIa) titration curves in the cases of normal blood composition, hemophilia A and B blood, blood lacking factor VII, blood lacking tissue factor pathway inhibitor, and diluted blood. We utilized the generated curves to perform systematic fold-change analyses for five quantitative parameters characterizing thrombin accumulation. RESULTS: The largest fold changes induced by increasing FVIIa concentration were observed for clotting time, thrombin peak time, and maximum slope of the thrombin curve. By contrast, thrombin peak height was much less affected by FVIIa titrations, and the area under the thrombin curve stayed practically unchanged. Comparisons with experimental data demonstrated that the computationally derived patterns can be observed in vitro. CONCLUSIONS: rFVIIa modulates thrombin generation primarily by accelerating the process, without significantly affecting the total amount of generated thrombin. link: http://identifiers.org/pubmed/21641634

Mitrophanov2013 - Simulation of Hockin Blood Coagulation Model under reduced temperature: MODEL1806250002v0.0.1

Blood coagulation model using an updated Hockin2002 model. New reactions for factor X and V activation by IXa and mIIa r…

Details

BACKGROUND: Hypothermia, which can result from tissue hypoperfusion, body exposure, and transfusion of cold resuscitation fluids, is a major factor contributing to coagulopathy of trauma and surgery. Despite considerable efforts, the mechanisms of hypothermia-induced blood coagulation impairment have not been fully understood. We introduce a kinetic modeling approach to investigate the effects of hypothermia on thrombin generation. METHODS: We extended a validated computational model to predict and analyze the impact of low temperatures (with or without concomitant blood dilution) on thrombin generation and its quantitative parameters. The computational model reflects the existing knowledge about the mechanistic details of thrombin generation biochemistry. We performed the analysis for an "average" subject, as well as for 472 subjects in the control group of the Leiden Thrombophilia Study. RESULTS: We computed and analyzed thousands of kinetic curves characterizing the generation of thrombin and the formation of the thrombin-antithrombin complex (TAT). In all simulations, hypothermia in the temperature interval 31°C to 36°C progressively slowed down thrombin generation, as reflected by clotting time, thrombin peak time, and prothrombin time, which increased in all subjects (P < 10(-5)). Maximum slope of the thrombin curve was progressively decreased, and the area under the thrombin curve was increased in hypothermia (P < 10(-5)); thrombin peak height remained practically unaffected. TAT formation was noticeably delayed (P < 10(-5)), but the final TAT levels were not significantly affected. Hypothermia-induced fold changes in the affected thrombin generation parameters were larger for lower temperatures, but were practically independent of the parameter itself and of the subjects' clotting factor composition, despite substantial variability in the subject group. Hypothermia and blood dilution acted additively on the thrombin generation parameters. CONCLUSIONS: We developed a general computational strategy that can be used to simulate the effects of changing temperature on the kinetics of biochemical systems and applied this strategy to analyze the effects of hypothermia on thrombin generation. We found that thrombin generation can be noticeably impaired in subjects with different blood plasma composition even in moderate hypothermia. Our work provides mechanistic support to the notion that thrombin generation impairment may be a key factor in coagulopathy induced by hypothermia and complicated by blood plasma dilution. link: http://identifiers.org/pubmed/23868891

Mitrophanov2014 - Extended Hockin Blood Coagulation Model with additional fibrin-associated species: MODEL1806260001v0.0.1

Mathematical model of the blood coagulation cascade. Extended Hockin model with contributions from Kim2007, Naski1991, S…

Details

Current mechanistic knowledge of protein interactions driving blood coagulation has come largely from experiments with simple synthetic systems, which only partially represent the molecular composition of human blood plasma. Here, we investigate the ability of the suggested molecular mechanisms to account for fibrin generation and degradation kinetics in diverse, physiologically relevant in vitro systems. We represented the protein interaction network responsible for thrombin generation, fibrin formation, and fibrinolysis as a computational kinetic model and benchmarked it against published and newly generated data reflecting diverse experimental conditions. We then applied the model to investigate the ability of fibrinogen and a recently proposed prothrombin complex concentrate composition, PCC-AT (a combination of the clotting factors II, IX, X, and antithrombin), to restore normal thrombin and fibrin generation in diluted plasma. The kinetic model captured essential features of empirically detected effects of prothrombin, fibrinogen, and thrombin-activatable fibrinolysis inhibitor titrations on fibrin formation and degradation kinetics. Moreover, the model qualitatively predicted the impact of tissue factor and tPA/tenecteplase level variations on the fibrin output. In the majority of considered cases, PCC-AT combined with fibrinogen accurately approximated both normal thrombin and fibrin generation in diluted plasma, which could not be accomplished by fibrinogen or PCC-AT acting alone. We conclude that a common network of protein interactions can account for key kinetic features characterizing fibrin accumulation and degradation in human blood plasma under diverse experimental conditions. Combined PCC-AT/fibrinogen supplementation is a promising strategy to reverse the deleterious effects of dilution-induced coagulopathy associated with traumatic bleeding. link: http://identifiers.org/pubmed/24958246

Mitrophanov2015 - Simulating extended Hockin Blood Coagulation Model under varied pH: BIOMD0000000951v0.0.1

Mathematical model of the blood coagulation cascade with new kinetic rates to simulate acidosis. Extended Hockin2002 mod…

Details

Acidosis, a frequent complication of trauma and complex surgery, results from tissue hypoperfusion and IV resuscitation with acidic fluids. While acidosis is known to inhibit the function of distinct enzymatic reactions, its cumulative effect on the blood coagulation system is not fully understood. Here, we use computational modeling to test the hypothesis that acidosis delays and reduces the amount of thrombin generation in human blood plasma. Moreover, we investigate the sensitivity of different thrombin generation parameters to acidosis, both at the individual and population level.We used a kinetic model to simulate and analyze the generation of thrombin and thrombin-antithrombin complexes (TAT), which were the end points of this study. Large groups of temporal thrombin and TAT trajectories were simulated and used to calculate quantitative parameters, such as clotting time (CT), thrombin peak time, maximum slope of the thrombin curve, thrombin peak height, area under the thrombin trajectory (AUC), and prothrombin time. The resulting samples of parameter values at different pH levels were compared to assess the acidosis-induced effects. To investigate intersubject variability, we parameterized the computational model using the data on clotting factor composition for 472 subjects from the Leiden Thrombophilia Study. To compare acidosis-induced relative parameter changes in individual ("virtual") subjects, we estimated the probabilities of relative change patterns by counting the pattern occurrences in our virtual subjects. Distribution overlaps for thrombin generation parameters at distinct pH levels were quantified using the Bhattacharyya coefficient.Acidosis in the range of pH 6.9 to 7.3 progressively increased CT, thrombin peak time, AUC, and prothrombin time, while decreasing maximum slope of the thrombin curve and thrombin peak height (P < 10). Acidosis delayed the onset and decreased the amount of TAT generation (P < 10). As a measure of intrasubject variability, maximum slope of the thrombin curve and CT displayed the largest and second-largest acidosis-induced relative changes, and AUC displayed the smallest relative changes among all thrombin generation parameters in our virtual subject group (1-sided 95% lower confidence limit on the fraction of subjects displaying the patterns, 0.99). As a measure of intersubject variability, the overlaps between the maximum slope of the thrombin curve distributions at acidotic pH levels with the maximum slope of the thrombin curve distribution at physiological pH level systematically exceeded analogous distribution overlaps for CT, thrombin peak time, and prothrombin time.Acidosis affected all quantitative parameters of thrombin and TAT generation. While maximum slope of the thrombin curve showed the highest sensitivity to acidosis at the individual-subject level, it may be outperformed by CT, thrombin peak time, and prothrombin time as an indicator of acidosis at the subject-group level. link: http://identifiers.org/pubmed/25839182

Parameters:

NameDescription
k07 = 23000.0Reaction: IIa + VII => IIa + VIIa, Rate Law: compartment_1*k07*IIa*VII
k29 = 103.0; k30 = 1.0E8Reaction: Xa_Va + II => Xa_Va_II, Rate Law: compartment_1*(k30*Xa_Va*II-k29*Xa_Va_II)
k27 = 0.2; k28 = 4.0E8Reaction: Xa + Va => Xa_Va, Rate Law: compartment_1*(k28*Xa*Va-k27*Xa_Va)
k21 = 1.0E8; k20 = 0.001Reaction: IXa_VIIIa + X => IXa_VIIIa_X, Rate Law: compartment_1*(k21*IXa_VIIIa*X-k20*IXa_VIIIa_X)
k26_0 = 2.0E7Reaction: IIa + V => IIa + Va, Rate Law: compartment_1*k26_0*IIa*V
k40 = 490.0Reaction: IXa + ATIII => IXa_ATIII, Rate Law: compartment_1*k40*IXa*ATIII
k37 = 5.0E7Reaction: TF_VIIa + Xa_TFPI => TF_VIIa_Xa_TFPI, Rate Law: compartment_1*k37*TF_VIIa*Xa_TFPI
k41 = 7100.0Reaction: IIa + ATIII => IIa_ATIII, Rate Law: compartment_1*k41*IIa*ATIII
k9 = 2.5E7; k8 = 1.05Reaction: TF_VIIa + X => TF_VIIa_X, Rate Law: compartment_1*(k9*TF_VIIa*X-k8*TF_VIIa_X)
k43_0 = 5700.0Reaction: IXa + X => IXa + Xa, Rate Law: compartment_1*k43_0*IXa*X
k25 = 0.001Reaction: IXa_VIIIa => VIIIa1_L + VIIIa2 + IXa, Rate Law: compartment_1*k25*IXa_VIIIa
k38 = 4200.0Reaction: Xa + ATIII => Xa_ATIII, Rate Law: compartment_1*k38*Xa*ATIII
k16_0 = 7500.0Reaction: Xa + II => Xa + IIa, Rate Law: compartment_1*k16_0*Xa*II
k23 = 22000.0; k24 = 0.006Reaction: VIIIa => VIIIa1_L + VIIIa2, Rate Law: compartment_1*(k24*VIIIa-k23*VIIIa1_L*VIIIa2)
k35 = 1.1E-4; k36 = 3.2E8Reaction: TF_VIIa_Xa + TFPI => TF_VIIa_Xa_TFPI, Rate Law: compartment_1*(k36*TF_VIIa_Xa*TFPI-k35*TF_VIIa_Xa_TFPI)
k22_0 = 8.2Reaction: IXa_VIIIa_X => IXa_VIIIa + Xa, Rate Law: compartment_1*k22_0*IXa_VIIIa_X
k44_0 = 3000000.0Reaction: mIIa + V => mIIa + Va, Rate Law: compartment_1*k44_0*mIIa*V
k32_0 = 2.3E8Reaction: mIIa + Xa_Va => IIa + Xa_Va, Rate Law: compartment_1*k32_0*mIIa*Xa_Va
k15_0 = 1.8Reaction: TF_VIIa_IX => TF_VIIa + IXa, Rate Law: compartment_1*k15_0*TF_VIIa_IX
k4 = 2.3E7; k3 = 0.0031Reaction: TF + VIIa => TF_VIIa, Rate Law: compartment_1*(k4*TF*VIIa-k3*TF_VIIa)
k39 = 7100.0Reaction: mIIa + ATIII => mIIa_ATIII, Rate Law: compartment_1*k39*mIIa*ATIII
k34 = 900000.0; k33 = 3.6E-4Reaction: Xa + TFPI => Xa_TFPI, Rate Law: compartment_1*(k34*Xa*TFPI-k33*Xa_TFPI)
k2 = 3200000.0; k1 = 0.0031Reaction: TF + VII => TF_VII, Rate Law: compartment_1*(k2*TF*VII-k1*TF_VII)
k10_0 = 6.0Reaction: TF_VIIa_X => TF_VIIa_Xa, Rate Law: compartment_1*k10_0*TF_VIIa_X
k12 = 2.2E7; k11 = 19.0Reaction: TF_VIIa + Xa => TF_VIIa_Xa, Rate Law: compartment_1*(k12*TF_VIIa*Xa-k11*TF_VIIa_Xa)
k42 = 230.0Reaction: TF_VIIa + ATIII => TF_VIIa_ATIII, Rate Law: compartment_1*k42*TF_VIIa*ATIII
k31_0 = 63.5Reaction: Xa_Va_II => Xa_Va + mIIa, Rate Law: compartment_1*k31_0*Xa_Va_II
k06 = 1.3E7Reaction: Xa + VII => Xa + VIIa, Rate Law: compartment_1*k06*Xa*VII
k05 = 440000.0Reaction: TF_VIIa + VII => TF_VIIa + VIIa, Rate Law: compartment_1*k05*TF_VIIa*VII
k19 = 1.0E7; k18 = 0.005Reaction: IXa + VIIIa => IXa_VIIIa, Rate Law: compartment_1*(k19*IXa*VIIIa-k18*IXa_VIIIa)
k14 = 1.0E7; k13 = 2.4Reaction: TF_VIIa + IX => TF_VIIa_IX, Rate Law: compartment_1*(k14*TF_VIIa*IX-k13*TF_VIIa_IX)
k17_0 = 2.0E7Reaction: IIa + VIII => IIa + VIIIa, Rate Law: compartment_1*k17_0*IIa*VIII

States:

NameDescription
IIa ATIII[Prothrombin; Antithrombin-III]
TFPI[Tissue factor pathway inhibitor]
Xa ATIII[Antithrombin-III; Coagulation factor X]
VIII[Coagulation factor VIII]
ATIII[Antithrombin-III]
Xa Va II[Coagulation factor V; Coagulation factor X; Prothrombin]
V[Coagulation factor V]
Xa[Coagulation factor X]
VIIIa1 L[Coagulation factor VIII]
TF VIIa ATIII[Coagulation factor VII; Antithrombin-III; Tissue factor]
IXa ATIII[Coagulation factor IX; Antithrombin-III]
TF VIIa X[Coagulation factor X; Coagulation factor VII; Tissue factor]
TF[Tissue factor]
TF VIIa Xa[Coagulation factor X; Coagulation factor VII; Tissue factor]
TF VIIa Xa TFPI[Tissue factor; Coagulation factor X; Coagulation factor VII; Tissue factor pathway inhibitor]
mIIa ATIII[Prothrombin; Antithrombin-III]
X[Coagulation factor X]
Xa Va[Coagulation factor X; Coagulation factor V]
TF VII[Coagulation factor VII; Tissue factor]
VIIIa2[Coagulation factor VIII]
TF VIIa[Tissue factor; Coagulation factor VII]
VIIIa[Coagulation factor VIII]
Va[Coagulation factor V]
mIIa[Prothrombin]
IIa[Prothrombin]
Xa TFPI[Coagulation factor X; Tissue factor pathway inhibitor]
VIIa[Coagulation factor VII]
IXa VIIIa X[Coagulation factor IX; Coagulation factor X; Coagulation factor VIII]
TF VIIa IX[Coagulation factor IX; Tissue factor; Coagulation factor VII]
IXa[Coagulation factor IX]
VII[Coagulation factor VII]
II[Prothrombin]
IX[Coagulation factor IX]
IXa VIIIa[Coagulation factor VIII; Coagulation factor IX]

Mitrophanov2016 - Extended Mitrophanov2011 Blood Coagulation Model (additional thrombin reactions): MODEL1806280001v0.0.1

Mathematical model of blood coagulation. Extended model of Mitrophanov2011 (which is an extension of Hockin2002). Additi…

Details

The use of prothrombin complex concentrates in trauma- and surgery-induced coagulopathy is complicated by the possibility of thromboembolic events. To explore the effects of these agents on thrombin generation (TG), we investigated combinations of coagulation factors equivalent to 3- and 4-factor prothrombin complex concentrates with and without added antithrombin (AT), as well as recombinant factor VIIa (rFVIIa), in a dilutional model. These data were then used to develop a computational model to test whether such a model could predict the TG profiles of these agents used to treat dilutional coagulopathy.We measured TG in plasma collected from 10 healthy volunteers using Calibrated Automated Thrombogram. TG measurements were performed in undiluted plasma, 3-fold saline-diluted plasma, and diluted plasma supplemented with the following factors: rFVIIa (group rFVIIa); factors (F)II, FIX, FX, and AT (group "combination of coagulation factors" [CCF]-AT); or FII, FVII, FIX, and FX (group CCF-FVII). We extended an existing computational model of TG to include additional reactions that impact the Calibrated Automated Thrombogram readout. We developed and applied a computational strategy to train the model using only a subset of the obtained TG data and used the remaining data for model validation.rFVIIa decreased lag time and the time to thrombin peak generation beyond their predilution levels (P < 0.001) but did not restore normal thrombin peak height (P < 0.001). CCF-FVII supplementation decreased lag time (P = 0.034) and thrombin peak time (P < 0.001) and increased both peak height (P < 0.001) and endogenous thrombin potential (P = 0.055) beyond their predilution levels. CCF-AT supplementation in diluted plasma resulted in an improvement in TG without causing the exaggerated effects of rFVIIa and CCF-FVII supplementation. The differences between the effects of CCF-AT and supplementation with rFVIIa and CCF-FVII were significant for lag time (P < 0.001 and P = 0.005, respectively), time to thrombin peak (P < 0.001 and P = 0.004, respectively), velocity index (P < 0.001 and P = 0.019, respectively), thrombin peak height (P < 0.001 for both comparisons), and endogenous thrombin potential (P = 0.034 and P = 0.019, respectively). The computational model generated subject-specific predictions and identified typical patterns of TG improvement.In this study of the effects of hemodilution, CCF-AT supplementation improved the dilution-impaired plasma TG potential in a more balanced way than either rFVIIa alone or CCF-FVII supplementation. Predictive computational modeling can guide plasma dilution/supplementation experiments. link: http://identifiers.org/pubmed/27541717

Mittler1998_HIV1_interactingTargetCells: MODEL1006230055v0.0.1

This a model from the article: Influence of delayed viral production on viral dynamics in HIV-1 infected patients. M…

Details

We present and analyze a model for the interaction of human immunodeficiency virus type 1 (HIV-1) with target cells that includes a time delay between initial infection and the formation of productively infected cells. Assuming that the variation among cells with respect to this 'intracellular' delay can be approximated by a gamma distribution, a high flexible distribution that can mimic a variety of biologically plausible delays, we provide analytical solutions for the expected decline in plasma virus concentration after the initiation of antiretroviral therapy with one or more protease inhibitors. We then use the model to investigate whether the parameters that characterize viral dynamics can be identified from biological data. Using non-linear least-squares regression to fit the model to simulated data in which the delays conform to a gamma distribution, we show that good estimates for free viral clearance rates, infected cell death rates, and parameters characterizing the gamma distribution can be obtained. For simulated data sets in which the delays were generated using other biologically plausible distributions, reasonably good estimates for viral clearance rates, infected cell death rates, and mean delay times can be obtained using the gamma-delay model. For simulated data sets that include added simulated noise, viral clearance rate estimates are not as reliable. If the mean intracellular delay is known, however, we show that reasonable estimates for the viral clearance rate can be obtained by taking the harmonic mean of viral clearance rate estimates from a group of patients. These results demonstrate that it is possible to incorporate distributed intracellular delays into existing models for HIV dynamics and to use these refined models to estimate the half-life of free virus from data on the decline in HIV-1 RNA following treatment. link: http://identifiers.org/pubmed/9780612

Mizuno2012 - AlzPathway: a comprehensive map of Alzheimer's disease: MODEL1504290001v0.0.1

Mizuno2012 - AlzPathway: a comprehensive map of Alzheimer's diseaseNon-kinetic molecular map. Pure SBML file of AlzPath…

Details

BACKGROUND: Alzheimer's disease (AD) is the most common cause of dementia among the elderly. To clarify pathogenesis of AD, thousands of reports have been accumulating. However, knowledge of signaling pathways in the field of AD has not been compiled as a database before. DESCRIPTION: Here, we have constructed a publicly available pathway map called "AlzPathway" that comprehensively catalogs signaling pathways in the field of AD. We have collected and manually curated over 100 review articles related to AD, and have built an AD pathway map using CellDesigner. AlzPathway is currently composed of 1347 molecules and 1070 reactions in neuron, brain blood barrier, presynaptic, postsynaptic, astrocyte, and microglial cells and their cellular localizations. AlzPathway is available as both the SBML (Systems Biology Markup Language) map for CellDesigner and the high resolution image map. AlzPathway is also available as a web service (online map) based on Payao system, a community-based, collaborative web service platform for pathway model curation, enabling continuous updates by AD researchers. CONCLUSIONS: AlzPathway is the first comprehensive map of intra, inter and extra cellular AD signaling pathways which can enable mechanistic deciphering of AD pathogenesis. The AlzPathway map is accessible at http://alzpathway.org/. link: http://identifiers.org/pubmed/22647208

Mkango2019 - Dynamics of Breast Cancer under DifferentRates of Chemoradiotherapy: MODEL1912120005v0.0.1

Dynamics of Breast Cancer under Different Rates of Chemoradiotherapy. Mkango SB1, Shaban N1, Mureithi E1, Ngoma T2. Auth…

Details

A type of cancer which originates from the breast tissue is referred to as breast cancer. Globally, it is the most common cause of death in women. Treatments such as radiotherapy, chemotherapy, hormone therapy, immunotherapy, and gene therapy are the main strategies in the fight against breast cancer. The present study aims at investigating the effects of the combined radiotherapy and chemotherapy as a way to treat breast cancer, and different treatment approaches are incorporated into the model. Also, the model is fitted to data on patients with breast cancer in Tanzania. We determine new treatment strategies, and finally, we show that when sufficient amount of chemotherapy and radiotherapy with a low decay rate is used, the drug will be significantly more effective in combating the disease while health cells remain above the threshold. link: http://identifiers.org/pubmed/31611927

Model of budding yeast critical cell size dependence on growth rate with parameter set 3114: MODEL1808310001v0.0.1

Model published in the paper Chaperone availability subordinates cell cycle entry to growth and stress by David F. Mo…

Details

The precise coordination of growth and proliferation has a universal prevalence in cell homeostasis. As a prominent property, cell size is modulated by the coordination between these processes in bacterial, yeast, and mammalian cells, but the underlying molecular mechanisms are largely unknown. Here, we show that multifunctional chaperone systems play a concerted and limiting role in cell-cycle entry, specifically driving nuclear accumulation of the G1 Cdk-cyclin complex. Based on these findings, we establish and test a molecular competition model that recapitulates cell-cycle-entry dependence on growth rate. As key predictions at a single-cell level, we show that availability of the Ydj1 chaperone and nuclear accumulation of the G1 cyclin Cln3 are inversely dependent on growth rate and readily respond to changes in protein synthesis and stress conditions that alter protein folding requirements. Thus, chaperone workload would subordinate Start to the biosynthetic machinery and dynamically adjust proliferation to the growth potential of the cell. link: http://identifiers.org/pubmed/30988162

MODEL1207300000_url.xml: MODEL1207300000v0.0.1

This model is from the article: A model of flux regulation in the cholesterol biosynthesis pathway: Immune mediated gr…

Details

The cholesterol biosynthesis pathway has recently been shown to play an important role in the innate immune response to viral infection with host protection occurring through a coordinate down regulation of the enzymes catalysing each metabolic step. In contrast, statin based drugs, which form the principle pharmaceutical agents for decreasing the activity of this pathway, target a single enzyme. Here, we build an ordinary differential equation model of the cholesterol biosynthesis pathway in order to investigate how the two regulatory strategies impact upon the behaviour of the pathway. We employ a modest set of assumptions: that the pathway operates away from saturation, that each metabolite is involved in multiple cellular interactions and that mRNA levels reflect enzyme concentrations. Using data taken from primary bone marrow derived macrophage cells infected with murine cytomegalovirus or treated with IFNγ, we show that, under these assumptions, coordinate down-regulation of enzyme activity imparts a graduated reduction in flux along the pathway. In contrast, modelling a statin-like treatment that achieves the same degree of down-regulation in cholesterol production, we show that this delivers a step change in flux along the pathway. The graduated reduction mediated by physiological coordinate regulation of multiple enzymes supports a mechanism that allows a greater level of specificity, altering cholesterol levels with less impact upon interactions branching from the pathway, than pharmacological step reductions. We argue that coordinate regulation is likely to show a long-term evolutionary advantage over single enzyme regulation. Finally, the results from our models have implications for future pharmaceutical therapies intended to target cholesterol production with greater specificity and fewer off target effects, suggesting that this can be achieved by mimicking the coordinated down-regulation observed in immunological responses. link: http://identifiers.org/pubmed/22664637

MODEL1703310000_url.xml: MODEL1703310000v0.0.1

Ramirez2017 - Human global metabolism in brown and white adipocytesRecon 2.1A, an update to Recon 2.1x, is suitable for…

Details

White adipocytes are specialized for energy storage, whereas brown adipocytes are specialized for energy expenditure. Explicating this difference can help identify therapeutic targets for obesity. A common tool to assess metabolic differences between such cells is the Seahorse Extracellular Flux (XF) Analyzer, which measures oxygen consumption and media acidification in the presence of different substrates and perturbagens. Here, we integrate the Analyzer's metabolic profile from human white and brown adipocytes with a genome-scale metabolic model to predict flux differences across the metabolic map. Predictions matched experimental data for the metabolite 4-aminobutyrate, the protein ABAT, and the fluxes for glucose, glutamine, and palmitate. We also uncovered a difference in how adipocytes dispose of nitrogenous waste, with brown adipocytes secreting less ammonia and more urea than white adipocytes. Thus, the method and software we developed allow for broader metabolic phenotyping and provide a distinct approach to uncovering metabolic differences. link: http://identifiers.org/pubmed/29241534

MODEL2021729243_url.xml: MODEL2021729243v0.0.1

This model originates from BioModels Database: A Database of Annotated Published Models (http://www.ebi.ac.uk/biomodels/…

Details

A better understanding of human metabolism and its relationship with diseases is an important task in human systems biology studies. In this paper, we present a high-quality human metabolic network manually reconstructed by integrating genome annotation information from different databases and metabolic reaction information from literature. The network contains nearly 3000 metabolic reactions, which were reorganized into about 70 human-specific metabolic pathways according to their functional relationships. By analysis of the functional connectivity of the metabolites in the network, the bow-tie structure, which was found previously by structure analysis, is reconfirmed. Furthermore, the distribution of the disease related genes in the network suggests that the IN (substrates) subset of the bow-tie structure has more flexibility than other parts. link: http://identifiers.org/pubmed/17882155

MODEL2021747594_url.xml: MODEL2021747594v0.0.1

This model originates from BioModels Database: A Database of Annotated Published Models (http://www.ebi.ac.uk/biomodels/…

Details

A better understanding of human metabolism and its relationship with diseases is an important task in human systems biology studies. In this paper, we present a high-quality human metabolic network manually reconstructed by integrating genome annotation information from different databases and metabolic reaction information from literature. The network contains nearly 3000 metabolic reactions, which were reorganized into about 70 human-specific metabolic pathways according to their functional relationships. By analysis of the functional connectivity of the metabolites in the network, the bow-tie structure, which was found previously by structure analysis, is reconfirmed. Furthermore, the distribution of the disease related genes in the network suggests that the IN (substrates) subset of the bow-tie structure has more flexibility than other parts. link: http://identifiers.org/pubmed/17882155

MODEL2426780967_url.xml: MODEL2426780967v0.0.1

**Increased glycolytic flux as an outcome of whole-genome duplication in yeast.** GC Conant and KH Wolfe, Mol Syst Biol…

Details

After whole-genome duplication (WGD), deletions return most loci to single copy. However, duplicate loci may survive through selection for increased dosage. Here, we show how the WGD increased copy number of some glycolytic genes could have conferred an almost immediate selective advantage to an ancestor of Saccharomyces cerevisiae, providing a rationale for the success of the WGD. We propose that the loss of other redundant genes throughout the genome resulted in incremental dosage increases for the surviving duplicated glycolytic genes. This increase gave post-WGD yeasts a growth advantage through rapid glucose fermentation; one of this lineage's many adaptations to glucose-rich environments. Our hypothesis is supported by data from enzyme kinetics and comparative genomics. Because changes in gene dosage follow directly from post-WGD deletions, dosage selection can confer an almost instantaneous benefit after WGD, unlike neofunctionalization or subfunctionalization, which require specific mutations. We also show theoretically that increased fermentative capacity is of greatest advantage when glucose resources are both large and dense, an observation potentially related to the appearance of angiosperms around the time of WGD. link: http://identifiers.org/pubmed/17667951

MODEL2427021978_url.xml: MODEL2427021978v0.0.1

**Increased glycolytic flux as an outcome of whole-genome duplication in yeast.** GC Conant and KH Wolfe, Mol Syst Biol…

Details

After whole-genome duplication (WGD), deletions return most loci to single copy. However, duplicate loci may survive through selection for increased dosage. Here, we show how the WGD increased copy number of some glycolytic genes could have conferred an almost immediate selective advantage to an ancestor of Saccharomyces cerevisiae, providing a rationale for the success of the WGD. We propose that the loss of other redundant genes throughout the genome resulted in incremental dosage increases for the surviving duplicated glycolytic genes. This increase gave post-WGD yeasts a growth advantage through rapid glucose fermentation; one of this lineage's many adaptations to glucose-rich environments. Our hypothesis is supported by data from enzyme kinetics and comparative genomics. Because changes in gene dosage follow directly from post-WGD deletions, dosage selection can confer an almost instantaneous benefit after WGD, unlike neofunctionalization or subfunctionalization, which require specific mutations. We also show theoretically that increased fermentative capacity is of greatest advantage when glucose resources are both large and dense, an observation potentially related to the appearance of angiosperms around the time of WGD. link: http://identifiers.org/pubmed/17667951

MODEL3631586579_url.xml: MODEL3631586579v0.0.1

This is a kinetic model of the monomeric photosystem II (PSII). The model is partially based on the earlier model use…

Details

MOTIVATION: It is a question of whether the supramolecular organization of the protein complex has an impact on its function, or not. In the case of the photosystem II (PSII), water splitting might be influenced by cooperation of the PSIIs. Since PSII is the source of the atmospheric oxygen and because better understanding of the water splitting may contribute to the effective use of water as an alternative energy source, possible cooperation should be analyzed and discussed. RESULTS: We suggest that the dimeric organization of the PSII induces cooperation in the water splitting. We show that the model of monomeric PSII is unable to produce the oxygen after the second short flash (associated with the double turnover of the PSII), in contrast to the experimental data and model of dimeric PSII with considered cooperation. On the basis of this fact and partially from the support from other studies, we concluded that the double turnover of the PSII induced by short flashes might be caused by the cooperation in the water splitting. We further discuss a possibility that the known pathway of the electron transport through the PSII might be incomplete and besides D1-Y161, other cofactor which is able to oxidize the special chlorophyll pair (P680) must be considered in the monomeric PSII to explain the oxygen production after the second short flash. AVAILABILITY: Commented SBML codes (.XML files) of the monomeric and dimeric PSII models will be available (at the time of publication) in the BioModels database (www.ebi.ac.uk/biomodels). link: http://identifiers.org/pubmed/18845578

MODEL3632127506_url.xml: MODEL3632127506v0.0.1

This is a kinetic model of the dimeric photosystem II (PSII). The model has partially based on the earlier model used…

Details

MOTIVATION: It is a question of whether the supramolecular organization of the protein complex has an impact on its function, or not. In the case of the photosystem II (PSII), water splitting might be influenced by cooperation of the PSIIs. Since PSII is the source of the atmospheric oxygen and because better understanding of the water splitting may contribute to the effective use of water as an alternative energy source, possible cooperation should be analyzed and discussed. RESULTS: We suggest that the dimeric organization of the PSII induces cooperation in the water splitting. We show that the model of monomeric PSII is unable to produce the oxygen after the second short flash (associated with the double turnover of the PSII), in contrast to the experimental data and model of dimeric PSII with considered cooperation. On the basis of this fact and partially from the support from other studies, we concluded that the double turnover of the PSII induced by short flashes might be caused by the cooperation in the water splitting. We further discuss a possibility that the known pathway of the electron transport through the PSII might be incomplete and besides D1-Y161, other cofactor which is able to oxidize the special chlorophyll pair (P680) must be considered in the monomeric PSII to explain the oxygen production after the second short flash. AVAILABILITY: Commented SBML codes (.XML files) of the monomeric and dimeric PSII models will be available (at the time of publication) in the BioModels database (www.ebi.ac.uk/biomodels). link: http://identifiers.org/pubmed/18845578

MODEL6624091635_url.xml: MODEL6624091635v0.0.1

This model originates from BioModels Database: A Database of Annotated Published Models (http://www.ebi.ac.uk/biomodels/…

Details

Glucose addition and subsequent run-out experiments were compared to simulations with a detailed glycolytic model of Lactococcus lactis. The model was constructed largely on bases of enzyme kinetic data taken from literature and not adjusted for the specific simulations shown here. Upon glucose depletion a rapid increase in PEP, inorganic phosphate and a gradual decrease in fructose 1,6-bisphosphate (FBP) were measured and predicted by simulation. The dynamic changes in these and other intermediate concentrations as measured in the experiments were well predicted by the kinetic model. link: http://identifiers.org/pubmed/12241048

MODEL8262229752_url.xml: MODEL8262229752v0.0.1

This model originates from BioModels Database: A Database of Annotated Published Models. It is copyright (c) 2005-2011 T…

Details

Quorum sensing (QS) is an important determinant of bacterial phenotype. Many cell functions are regulated by intricate and multimodal QS signal transduction processes. The LuxS/AI-2 QS system is highly conserved among Eubacteria and AI-2 is reported as a 'universal' signal molecule. To understand the hierarchical organization of AI-2 circuitry, a comprehensive approach incorporating stochastic simulations was developed. We investigated the synthesis, uptake, and regulation of AI-2, developed testable hypotheses, and made several discoveries: (1) the mRNA transcript and protein levels of AI-2 synthases, Pfs and LuxS, do not contribute to the dramatically increased level of AI-2 found when cells are grown in the presence of glucose; (2) a concomitant increase in metabolic flux through this synthesis pathway in the presence of glucose only partially accounts for this difference. We predict that 'high-flux' alternative pathways or additional biological steps are involved in AI-2 synthesis; and (3) experimental results validate this hypothesis. This work demonstrates the utility of linking cell physiology with systems-based stochastic models that can be assembled de novo with partial knowledge of biochemical pathways. link: http://identifiers.org/pubmed/17170762

Mol2013 - Immune Signal Transduction in Leishmaniasis: BIOMD0000000477v0.0.1

Created by The MathWorks, Inc. SimBiology tool, Version 3.3

Details

Modulated immune signal (CD14-TLR and TNF) in leishmaniasis can be linked to EGFR pathway involved in wound healing, through crosstalk points. This signaling network can be further linked to a synthetic gene circuit acting as a positive feedback loop to elicit a synchronized intercellular communication among the immune cells which may contribute to a better understanding of signaling dynamics in leishmaniasis.Network reconstruction with positive feedback loop, simulation (ODE 15s solver) and sensitivity analysis of CD14-TLR, TNF and EGFR was done in SimBiology (MATLAB 7.11.1). Cytoscape and adjacency matrix were used to calculate network topology. PCA was extracted by using sensitivity coefficient in MATLAB. Model reduction was done using time, flux and sensitivity score.Network has five crosstalk points: NIK, IκB-NFκB and MKK (4/7, 3/6, 1/2) which show high flux and sensitivity. PI3K in EGFR pathway shows high flux and sensitivity. PCA score was high for cytoplasmic ERK1/2, PI3K, Atk, STAT1/3 and nuclear JNK. Of the 125 parameters, 20% are crucial as deduced by model reduction.EGFR can be linked to CD14-TLR and TNF through the MAPK crosstalk points. These pathways may be controlled through Ras and Raf that lie upstream of signaling components ERK ½ (c) and JNK (n) that have a high PCA score via a synthetic gene circuit for activating cell-cell communication to elicit an inflammatory response. Also a disease resolving effect may be achieved through PI3K in the EGFR pathway.The reconstructed signaling network can be linked to a gene circuit with a positive feedback loop, for cell-cell communication resulting in synchronized response in the immune cell population, for disease resolving effect in leishmaniasis. link: http://identifiers.org/pubmed/23994140

Parameters:

NameDescription
mw286a7792_09c4_443e_98f4_a68f66a1f380=1.0 1/secondReaction: mwb71eb539_dca6_47ab_8df5_430d84af0bfb => mw97345a67_a8e8_42aa_8e62_69e9d2b6cf45; mwb71eb539_dca6_47ab_8df5_430d84af0bfb, Rate Law: mw286a7792_09c4_443e_98f4_a68f66a1f380*mwb71eb539_dca6_47ab_8df5_430d84af0bfb
mwf89fc9a4_ad1e_4e59_8a06_4b8dc2cc84a7=0.28 nanomole/second; mwaad66a38_26d2_41fc_9261_79c57500a6d4=1.5 nanomoleReaction: mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa => mwf8cfed1b_6fcf_4cba_bc30_b44490814a7a; mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa, Rate Law: mwf89fc9a4_ad1e_4e59_8a06_4b8dc2cc84a7*mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa/(mwaad66a38_26d2_41fc_9261_79c57500a6d4+mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa)
mwb38e4258_82d9_4b48_8059_eccf9fd6f8e3=0.2 nanomole; mw99befd62_975f_49e1_bfaf_22a482ce44ea=1.2 nanomole/secondReaction: mwb4633da9_f9d6_4ad8_a7e5_da075c830e17 => mw7204ab72_2ee5_4b92_b420_2583dacc4343; mwb4633da9_f9d6_4ad8_a7e5_da075c830e17, Rate Law: mw99befd62_975f_49e1_bfaf_22a482ce44ea*mwb4633da9_f9d6_4ad8_a7e5_da075c830e17/(mwb38e4258_82d9_4b48_8059_eccf9fd6f8e3+mwb4633da9_f9d6_4ad8_a7e5_da075c830e17)
mw9d622ba3_b43b_4101_bef8_c964c2f158a0=0.99 nanomole/second; mw2b1ea101_d4a1_42e9_a70f_cb8026911ed5=0.08 nanomoleReaction: mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa => mw805b55df_cc91_4227_bb52_930e961b682c; mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa, Rate Law: mw9d622ba3_b43b_4101_bef8_c964c2f158a0*mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa/(mw2b1ea101_d4a1_42e9_a70f_cb8026911ed5+mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa)
mweaee0b65_7c40_4c9e_bd70_c5454eeb41fa=0.9 nanomole/second; mw84020ddc_e419_4aa4_ab12_e84989ad461d=0.3 nanomoleReaction: mw05469f51_73f7_4ba1_9f1a_bce5fea143c2 => mwf20834c8_a115_460b_859c_4e3ca1ffd953; mw05469f51_73f7_4ba1_9f1a_bce5fea143c2, Rate Law: mweaee0b65_7c40_4c9e_bd70_c5454eeb41fa*mw05469f51_73f7_4ba1_9f1a_bce5fea143c2/(mw84020ddc_e419_4aa4_ab12_e84989ad461d+mw05469f51_73f7_4ba1_9f1a_bce5fea143c2)
mw1c3fcb1f_0b90_46dd_b13a_2950fb9e18ae=0.2 nanomole/second; mw8a65d230_2abb_478d_ab8a_6719d972483d=0.01 nanomoleReaction: mw308b75ec_28b7_4d97_92e2_51a8ce04116a => mw75377e12_e23d_44b3_9823_5fac9b23edc8; mw308b75ec_28b7_4d97_92e2_51a8ce04116a, Rate Law: mw1c3fcb1f_0b90_46dd_b13a_2950fb9e18ae*mw308b75ec_28b7_4d97_92e2_51a8ce04116a/(mw8a65d230_2abb_478d_ab8a_6719d972483d+mw308b75ec_28b7_4d97_92e2_51a8ce04116a)
mw1670fb0f_e301_4b7a_93d4_35fe7f504e92=0.03 nanomole; mwcad6928f_259d_4125_987e_977e0c40ef7d=1.5 nanomole/secondReaction: mw46ee629a_dd6b_4163_9da1_2614bb1d74bc => mwb71eb539_dca6_47ab_8df5_430d84af0bfb; mw46ee629a_dd6b_4163_9da1_2614bb1d74bc, Rate Law: mwcad6928f_259d_4125_987e_977e0c40ef7d*mw46ee629a_dd6b_4163_9da1_2614bb1d74bc/(mw1670fb0f_e301_4b7a_93d4_35fe7f504e92+mw46ee629a_dd6b_4163_9da1_2614bb1d74bc)
mw107b07de_5145_436d_9fd7_e4e2103106d7=1.2 nanomole/second; mwd51a525a_5fea_42c6_a8fd_40429ee627cf=0.02 nanomoleReaction: mw805b55df_cc91_4227_bb52_930e961b682c => mw46ee629a_dd6b_4163_9da1_2614bb1d74bc; mw805b55df_cc91_4227_bb52_930e961b682c, Rate Law: mw107b07de_5145_436d_9fd7_e4e2103106d7*mw805b55df_cc91_4227_bb52_930e961b682c/(mwd51a525a_5fea_42c6_a8fd_40429ee627cf+mw805b55df_cc91_4227_bb52_930e961b682c)
mwd4bfd4cc_6fd4_4b3b_980d_547ce2740b7e=2.2 nanomole/second; mw9ac53fed_0388_4261_b457_030cd631fa0e=0.1 nanomoleReaction: mw7204ab72_2ee5_4b92_b420_2583dacc4343 => mw6939cefe_e7ff_4a3f_b45b_a9234d1b5573; mw7204ab72_2ee5_4b92_b420_2583dacc4343, Rate Law: mwd4bfd4cc_6fd4_4b3b_980d_547ce2740b7e*mw7204ab72_2ee5_4b92_b420_2583dacc4343/(mw9ac53fed_0388_4261_b457_030cd631fa0e+mw7204ab72_2ee5_4b92_b420_2583dacc4343)
mw0e1c63a9_8b8a_4ec7_9608_0059208d992f=0.01 nanomole; mw6ac279a2_23fe_4e48_a910_2a94ef61244c=0.1 nanomole/secondReaction: mw308b75ec_28b7_4d97_92e2_51a8ce04116a => mw46ee629a_dd6b_4163_9da1_2614bb1d74bc; mw308b75ec_28b7_4d97_92e2_51a8ce04116a, Rate Law: mw6ac279a2_23fe_4e48_a910_2a94ef61244c*mw308b75ec_28b7_4d97_92e2_51a8ce04116a/(mw0e1c63a9_8b8a_4ec7_9608_0059208d992f+mw308b75ec_28b7_4d97_92e2_51a8ce04116a)
mw4c0ee457_fb1c_48fa_a0b7_ff10d632d1e0=2.0 1/secondReaction: mw4079e13c_446e_4aa2_9ec4_233583833d02 => mwe5fade7d_1715_4bb1_843f_923da8ecddf1; mw4079e13c_446e_4aa2_9ec4_233583833d02, Rate Law: mw4c0ee457_fb1c_48fa_a0b7_ff10d632d1e0*mw4079e13c_446e_4aa2_9ec4_233583833d02
mw9d566811_669e_4b95_8452_c4853f54a2de=0.35 1/secondReaction: mwf20834c8_a115_460b_859c_4e3ca1ffd953 => mwb4633da9_f9d6_4ad8_a7e5_da075c830e17; mwf20834c8_a115_460b_859c_4e3ca1ffd953, Rate Law: mw9d566811_669e_4b95_8452_c4853f54a2de*mwf20834c8_a115_460b_859c_4e3ca1ffd953
mwc29ba5b1_b0e7_4fa1_9e46_a4c0bdbdacc4=0.2 nanomole; mw5990b7f9_7d15_4306_9047_6237ecf066ca=0.9 nanomole/secondReaction: mw75377e12_e23d_44b3_9823_5fac9b23edc8 => mw67d0cf04_d6a7_4725_a869_098a96a3350d; mw75377e12_e23d_44b3_9823_5fac9b23edc8, Rate Law: mw5990b7f9_7d15_4306_9047_6237ecf066ca*mw75377e12_e23d_44b3_9823_5fac9b23edc8/(mwc29ba5b1_b0e7_4fa1_9e46_a4c0bdbdacc4+mw75377e12_e23d_44b3_9823_5fac9b23edc8)
mwa68f7af3_30af_4fa0_9290_9e005c875763=1.4 1/secondReaction: mw6939cefe_e7ff_4a3f_b45b_a9234d1b5573 => mw8a358487_b18b_42df_a646_cd75eb5bfcc2; mw6939cefe_e7ff_4a3f_b45b_a9234d1b5573, Rate Law: mwa68f7af3_30af_4fa0_9290_9e005c875763*mw6939cefe_e7ff_4a3f_b45b_a9234d1b5573
mw8adff9cb_4657_413f_a2bd_100d4aa53076=0.98 nanomole/second; mwc9cf88fa_c525_4372_80e1_c72b1cc758f1=0.15 nanomoleReaction: mw702be69a_eb4f_425e_87c7_ef7d85254536 => mwbee11634_55df_4a3f_998a_634dfaf46fd7; mw702be69a_eb4f_425e_87c7_ef7d85254536, Rate Law: mw8adff9cb_4657_413f_a2bd_100d4aa53076*mw702be69a_eb4f_425e_87c7_ef7d85254536/(mwc9cf88fa_c525_4372_80e1_c72b1cc758f1+mw702be69a_eb4f_425e_87c7_ef7d85254536)
mw1a4dcdaf_ff4b_41a9_ac1d_79fd2d942260=0.6 1/secondReaction: mw0be0d193_fd6b_4824_8928_dbade8b5c99c => mw280197c8_98de_43f0_bf01_0f332a1ab689; mw0be0d193_fd6b_4824_8928_dbade8b5c99c, Rate Law: mw1a4dcdaf_ff4b_41a9_ac1d_79fd2d942260*mw0be0d193_fd6b_4824_8928_dbade8b5c99c
mw85a8c1da_f29f_4dcf_a515_bf9f9921240b=0.2 nanomole; mwf5a1613b_fb22_43b0_b95a_2c18ecbcedd8=0.5 nanomole/secondReaction: mw308b75ec_28b7_4d97_92e2_51a8ce04116a => mw136c8391_14f4_4a28_83a3_35cc74a2e040; mw308b75ec_28b7_4d97_92e2_51a8ce04116a, Rate Law: mwf5a1613b_fb22_43b0_b95a_2c18ecbcedd8*mw308b75ec_28b7_4d97_92e2_51a8ce04116a/(mw85a8c1da_f29f_4dcf_a515_bf9f9921240b+mw308b75ec_28b7_4d97_92e2_51a8ce04116a)
mwa4c28075_8524_4874_aee5_c38231bfbaae=1.5 nanomole; mw66285193_607e_42b6_b726_c2409a2ce563=1.0 nanomole/secondReaction: mw3832f277_aef2_4f1d_87af_abc2a3c1a7d5 => mw13651143_feb5_49a5_adab_9105c2647446; mw3832f277_aef2_4f1d_87af_abc2a3c1a7d5, Rate Law: mw66285193_607e_42b6_b726_c2409a2ce563*mw3832f277_aef2_4f1d_87af_abc2a3c1a7d5/(mwa4c28075_8524_4874_aee5_c38231bfbaae+mw3832f277_aef2_4f1d_87af_abc2a3c1a7d5)
mw6a74caa7_9d44_449b_854b_c1678b36ac1d=0.8 nanomole; mw78df1f4c_2a96_4d8f_a009_c19ba0ec406a=0.2 nanomole/secondReaction: mw280197c8_98de_43f0_bf01_0f332a1ab689 => mw9bb804c9_3e4e_4684_9f6b_4e6f6706a58e; mw280197c8_98de_43f0_bf01_0f332a1ab689, Rate Law: mw78df1f4c_2a96_4d8f_a009_c19ba0ec406a*mw280197c8_98de_43f0_bf01_0f332a1ab689/(mw6a74caa7_9d44_449b_854b_c1678b36ac1d+mw280197c8_98de_43f0_bf01_0f332a1ab689)
mw9a480703_d4bb_4de8_8975_13a18205ce53=0.19 1/secondReaction: mw13651143_feb5_49a5_adab_9105c2647446 => mw17ae9adc_54ab_407b_a34d_8413a3a10cc6; mw13651143_feb5_49a5_adab_9105c2647446, Rate Law: mw9a480703_d4bb_4de8_8975_13a18205ce53*mw13651143_feb5_49a5_adab_9105c2647446
mw661c7759_2bd3_4c93_bb0a_823bb37b9820=0.299 1/secondReaction: mw280197c8_98de_43f0_bf01_0f332a1ab689 => mw323a57b4_8e59_4116_9ad1_fe547b89c858; mw280197c8_98de_43f0_bf01_0f332a1ab689, Rate Law: mw661c7759_2bd3_4c93_bb0a_823bb37b9820*mw280197c8_98de_43f0_bf01_0f332a1ab689
mwa7160f91_3c68_402a_b3bd_acd8490c5d2d=0.56 nanomole; mw2b6193d2_d588_46b7_8463_ce7bc30e1575=1.3 nanomole/secondReaction: mw2dc73059_a841_48d5_b4bd_3ac24d94c42e => mw7204ab72_2ee5_4b92_b420_2583dacc4343; mw2dc73059_a841_48d5_b4bd_3ac24d94c42e, Rate Law: mw2b6193d2_d588_46b7_8463_ce7bc30e1575*mw2dc73059_a841_48d5_b4bd_3ac24d94c42e/(mwa7160f91_3c68_402a_b3bd_acd8490c5d2d+mw2dc73059_a841_48d5_b4bd_3ac24d94c42e)
mw5aa11378_86b4_45f6_aea1_27208e47e559=0.72 1/secondReaction: mwe5fade7d_1715_4bb1_843f_923da8ecddf1 => mw262497ec_3d54_4367_bfe3_76a9c57497cb; mwe5fade7d_1715_4bb1_843f_923da8ecddf1, Rate Law: mw5aa11378_86b4_45f6_aea1_27208e47e559*mwe5fade7d_1715_4bb1_843f_923da8ecddf1
mw251bb80a_5527_4b9c_9834_99556d4e824a=0.01 nanomole; mw75017b10_387d_43e4_9fb1_fed7ce6bd490=0.3 nanomole/secondReaction: mw308b75ec_28b7_4d97_92e2_51a8ce04116a => mw702be69a_eb4f_425e_87c7_ef7d85254536; mw308b75ec_28b7_4d97_92e2_51a8ce04116a, Rate Law: mw75017b10_387d_43e4_9fb1_fed7ce6bd490*mw308b75ec_28b7_4d97_92e2_51a8ce04116a/(mw251bb80a_5527_4b9c_9834_99556d4e824a+mw308b75ec_28b7_4d97_92e2_51a8ce04116a)
mw2b132eeb_ce2a_4a53_8c22_c102ebd2edb9=1.1 1/secondReaction: mwc844b7c0_98f5_4d0d_8f0c_00dfe8b54e6d => mw4d2e70a7_f499_461d_ae18_bc53b365b091; mwc844b7c0_98f5_4d0d_8f0c_00dfe8b54e6d, Rate Law: mw2b132eeb_ce2a_4a53_8c22_c102ebd2edb9*mwc844b7c0_98f5_4d0d_8f0c_00dfe8b54e6d
mw0733f43b_b430_40c4_8b93_1555a4bdbaa1=0.6 nanomole/second; mw56211dd8_6a88_465e_bed2_f603bf8c5b52=0.6 nanomoleReaction: mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa => mw136c8391_14f4_4a28_83a3_35cc74a2e040; mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa, Rate Law: mw0733f43b_b430_40c4_8b93_1555a4bdbaa1*mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa/(mw56211dd8_6a88_465e_bed2_f603bf8c5b52+mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa)
mw6834a7ac_63c4_4741_b0fc_069c665f1de2=1.18 1/secondReaction: mw8cc67de0_64e6_428f_ab09_4c2825cc172c => mw6ee00a71_ab68_454b_b1cd_60c1ebd19cfa; mw8cc67de0_64e6_428f_ab09_4c2825cc172c, Rate Law: mw6834a7ac_63c4_4741_b0fc_069c665f1de2*mw8cc67de0_64e6_428f_ab09_4c2825cc172c
mwbad3f510_fbca_4aa7_a4c2_5c1b47297802=2.35 nanomole/second; mw2fa0d3fe_4e99_49d2_a339_089198589a1e=0.43 nanomole; mw4d5fd70d_8603_4056_adfa_5af26d657455=1.0Reaction: mw8a358487_b18b_42df_a646_cd75eb5bfcc2 => mwc844b7c0_98f5_4d0d_8f0c_00dfe8b54e6d; mw8a358487_b18b_42df_a646_cd75eb5bfcc2, Rate Law: mwbad3f510_fbca_4aa7_a4c2_5c1b47297802*mw8a358487_b18b_42df_a646_cd75eb5bfcc2^mw4d5fd70d_8603_4056_adfa_5af26d657455/(mw2fa0d3fe_4e99_49d2_a339_089198589a1e+mw8a358487_b18b_42df_a646_cd75eb5bfcc2^mw4d5fd70d_8603_4056_adfa_5af26d657455)
mwbf5d43e3_e386_4b05_997d_4e70cbff9498=0.15 nanomole; mw5fdc2487_13a9_449a_b90c_95446ddf7f37=1.3 nanomole/secondReaction: mw136c8391_14f4_4a28_83a3_35cc74a2e040 => mw2dc73059_a841_48d5_b4bd_3ac24d94c42e; mw136c8391_14f4_4a28_83a3_35cc74a2e040, Rate Law: mw5fdc2487_13a9_449a_b90c_95446ddf7f37*mw136c8391_14f4_4a28_83a3_35cc74a2e040/(mwbf5d43e3_e386_4b05_997d_4e70cbff9498+mw136c8391_14f4_4a28_83a3_35cc74a2e040)
mw883852ed_c433_4dec_baa0_386309fc085c=0.24 nanomole/second; mw6069097b_159a_4bcf_a591_e496d06cf0a9=1.2 nanomoleReaction: mw323a57b4_8e59_4116_9ad1_fe547b89c858 => mw173d8585_5817_4b4c_932a_cf7d673680ac; mw323a57b4_8e59_4116_9ad1_fe547b89c858, Rate Law: mw883852ed_c433_4dec_baa0_386309fc085c*mw323a57b4_8e59_4116_9ad1_fe547b89c858/(mw6069097b_159a_4bcf_a591_e496d06cf0a9+mw323a57b4_8e59_4116_9ad1_fe547b89c858)
mwd2f6a3b7_5a74_4d77_b40c_1a6713b98554=0.42 1/secondReaction: mw67d0cf04_d6a7_4725_a869_098a96a3350d => mw1f12e5bc_ebbc_4347_b6b7_5cd1740ac69a; mw67d0cf04_d6a7_4725_a869_098a96a3350d, Rate Law: mwd2f6a3b7_5a74_4d77_b40c_1a6713b98554*mw67d0cf04_d6a7_4725_a869_098a96a3350d
mw26de6022_cc14_484b_a172_db4173a1ccaa=0.7 nanomole/second; mw7e75e47c_6d88_49fb_a9c4_9154f12cc4d5=1.5 nanomoleReaction: mw32c21c39_237b_4d4c_bb5d_117cb30ce68a => mw75377e12_e23d_44b3_9823_5fac9b23edc8; mw32c21c39_237b_4d4c_bb5d_117cb30ce68a, Rate Law: mw26de6022_cc14_484b_a172_db4173a1ccaa*mw32c21c39_237b_4d4c_bb5d_117cb30ce68a/(mw7e75e47c_6d88_49fb_a9c4_9154f12cc4d5+mw32c21c39_237b_4d4c_bb5d_117cb30ce68a)
mwafa60fbe_9272_468d_94e7_b82b985f938c=0.61 1/secondReaction: mwbee11634_55df_4a3f_998a_634dfaf46fd7 => mwd9e7a9b9_6f1b_4bbc_afa5_6cb192b62ce8; mwbee11634_55df_4a3f_998a_634dfaf46fd7, Rate Law: mwafa60fbe_9272_468d_94e7_b82b985f938c*mwbee11634_55df_4a3f_998a_634dfaf46fd7
mw244e346b_4442_45db_864e_0442ceca94d1=0.5 nanomole; mwb3751ef8_2226_4ec3_9ac9_f92f5771a1a4=2.0E-4 nanomole/secondReaction: mw280197c8_98de_43f0_bf01_0f332a1ab689 => mw9a5baf6d_0285_4ad3_9499_059c553d9cf6; mw280197c8_98de_43f0_bf01_0f332a1ab689, Rate Law: mwb3751ef8_2226_4ec3_9ac9_f92f5771a1a4*mw280197c8_98de_43f0_bf01_0f332a1ab689/(mw244e346b_4442_45db_864e_0442ceca94d1+mw280197c8_98de_43f0_bf01_0f332a1ab689)
mw6e048357_d06d_4522_bb79_a91c4f53bda7=0.6 1/secondReaction: mwa5d6f7e4_dc4d_4931_91ce_1e78e7b2f195 => mw4079e13c_446e_4aa2_9ec4_233583833d02; mwa5d6f7e4_dc4d_4931_91ce_1e78e7b2f195, Rate Law: mw6e048357_d06d_4522_bb79_a91c4f53bda7*mwa5d6f7e4_dc4d_4931_91ce_1e78e7b2f195
mwf88d190e_a505_4f7e_ac8d_e43997c74b9c=0.62 nanomole; mw1a1570ff_e786_473f_860b_2e7694acfcc2=1.14 nanomole/secondReaction: mwd9e7a9b9_6f1b_4bbc_afa5_6cb192b62ce8 => mwfed5a135_c91b_4d20_91b2_3a61723544dd; mwd9e7a9b9_6f1b_4bbc_afa5_6cb192b62ce8, Rate Law: mw1a1570ff_e786_473f_860b_2e7694acfcc2*mwd9e7a9b9_6f1b_4bbc_afa5_6cb192b62ce8/(mwf88d190e_a505_4f7e_ac8d_e43997c74b9c+mwd9e7a9b9_6f1b_4bbc_afa5_6cb192b62ce8)
mw36ee8f87_d06f_4d16_ac13_f4075b56c6f4=0.55 1/secondReaction: mw262497ec_3d54_4367_bfe3_76a9c57497cb => mw8bffd47e_34de_4738_81bf_7a39a40b3ae8; mw262497ec_3d54_4367_bfe3_76a9c57497cb, Rate Law: mw36ee8f87_d06f_4d16_ac13_f4075b56c6f4*mw262497ec_3d54_4367_bfe3_76a9c57497cb
mw37ac6d2c_1be9_4998_a9c5_8761d3e0ba0f=0.2 nanomole; mw31c3bf7d_10cd_412a_9a76_0fb66845c18d=0.6 nanomole/secondReaction: mwf8cfed1b_6fcf_4cba_bc30_b44490814a7a => mw702be69a_eb4f_425e_87c7_ef7d85254536; mwf8cfed1b_6fcf_4cba_bc30_b44490814a7a, Rate Law: mw31c3bf7d_10cd_412a_9a76_0fb66845c18d*mwf8cfed1b_6fcf_4cba_bc30_b44490814a7a/(mw37ac6d2c_1be9_4998_a9c5_8761d3e0ba0f+mwf8cfed1b_6fcf_4cba_bc30_b44490814a7a)
mw4945db3d_e20c_4870_b96b_6fb98c4b12f6=1.0 nanomole; mwdeab2870_570e_4b2c_b73d_84c1ad8c2262=1.0 nanomole/secondReaction: mwb4633da9_f9d6_4ad8_a7e5_da075c830e17 => mw173d8585_5817_4b4c_932a_cf7d673680ac; mwb4633da9_f9d6_4ad8_a7e5_da075c830e17, Rate Law: mwdeab2870_570e_4b2c_b73d_84c1ad8c2262*mwb4633da9_f9d6_4ad8_a7e5_da075c830e17/(mw4945db3d_e20c_4870_b96b_6fb98c4b12f6+mwb4633da9_f9d6_4ad8_a7e5_da075c830e17)
mw78a1e67e_883c_497f_86a6_f85da783010e=0.2 nanomole/second; mw5d6cf9c6_4dc0_4fe6_9afc_da397fe896b2=1.5 nanomoleReaction: mw173d8585_5817_4b4c_932a_cf7d673680ac => mw32c21c39_237b_4d4c_bb5d_117cb30ce68a; mw173d8585_5817_4b4c_932a_cf7d673680ac, Rate Law: mw78a1e67e_883c_497f_86a6_f85da783010e*mw173d8585_5817_4b4c_932a_cf7d673680ac/(mw5d6cf9c6_4dc0_4fe6_9afc_da397fe896b2+mw173d8585_5817_4b4c_932a_cf7d673680ac)
mw0b0869f4_26bb_4d13_9124_b2c1b28e3ae1=1.5 nanomole; mw2f1f65d1_5633_4625_b2b7_0eb267eac293=0.4 nanomole/secondReaction: mw173d8585_5817_4b4c_932a_cf7d673680ac => mw702be69a_eb4f_425e_87c7_ef7d85254536; mw173d8585_5817_4b4c_932a_cf7d673680ac, Rate Law: mw2f1f65d1_5633_4625_b2b7_0eb267eac293*mw173d8585_5817_4b4c_932a_cf7d673680ac/(mw0b0869f4_26bb_4d13_9124_b2c1b28e3ae1+mw173d8585_5817_4b4c_932a_cf7d673680ac)
mw826aae9f_9728_4bbb_a11b_60578912218b=1.2 1/secondReaction: mw4d2e70a7_f499_461d_ae18_bc53b365b091 => mw8cc67de0_64e6_428f_ab09_4c2825cc172c; mw4d2e70a7_f499_461d_ae18_bc53b365b091, Rate Law: mw826aae9f_9728_4bbb_a11b_60578912218b*mw4d2e70a7_f499_461d_ae18_bc53b365b091
mw13b39522_0751_4041_a78e_871cd5d81592=0.2 nanomole; mw2a0659f9_eab8_4ada_8f82_23068b9986eb=1.5 nanomole/secondReaction: mw97345a67_a8e8_42aa_8e62_69e9d2b6cf45 => mw5c67812a_17f5_43cf_8acb_9bde272c1911; mw97345a67_a8e8_42aa_8e62_69e9d2b6cf45, Rate Law: mw2a0659f9_eab8_4ada_8f82_23068b9986eb*mw97345a67_a8e8_42aa_8e62_69e9d2b6cf45/(mw13b39522_0751_4041_a78e_871cd5d81592+mw97345a67_a8e8_42aa_8e62_69e9d2b6cf45)
mwb69d510c_dcde_4bfb_9e4a_89954f6a7bf5=1.5 nanomole; mw3690266b_c916_4ba1_a98a_b589dc75c1cd=0.8 nanomole/secondReaction: mw9bb804c9_3e4e_4684_9f6b_4e6f6706a58e => mw64453fc5_a275_4bba_84f0_2af249b31514; mw9bb804c9_3e4e_4684_9f6b_4e6f6706a58e, Rate Law: mw3690266b_c916_4ba1_a98a_b589dc75c1cd*mw9bb804c9_3e4e_4684_9f6b_4e6f6706a58e/(mwb69d510c_dcde_4bfb_9e4a_89954f6a7bf5+mw9bb804c9_3e4e_4684_9f6b_4e6f6706a58e)
mw933afd80_4eff_4c6c_967b_d15b2244e55d=1.0 nanomole; mw0f1ee85e_95a3_42c7_94ae_71f36061aaf0=1.0 nanomole/secondReaction: mwb4633da9_f9d6_4ad8_a7e5_da075c830e17 => mw32c21c39_237b_4d4c_bb5d_117cb30ce68a; mwb4633da9_f9d6_4ad8_a7e5_da075c830e17, Rate Law: mw0f1ee85e_95a3_42c7_94ae_71f36061aaf0*mwb4633da9_f9d6_4ad8_a7e5_da075c830e17/(mw933afd80_4eff_4c6c_967b_d15b2244e55d+mwb4633da9_f9d6_4ad8_a7e5_da075c830e17)
mw6d4dc2a5_6fe8_4d80_93f4_b9f438b6eb0e=1.0 nanomole; mw18e075a4_dde4_42be_9315_e0e90d461b99=0.6 nanomole/secondReaction: mw9a5baf6d_0285_4ad3_9499_059c553d9cf6 => mw05469f51_73f7_4ba1_9f1a_bce5fea143c2; mw9a5baf6d_0285_4ad3_9499_059c553d9cf6, Rate Law: mw18e075a4_dde4_42be_9315_e0e90d461b99*mw9a5baf6d_0285_4ad3_9499_059c553d9cf6/(mw6d4dc2a5_6fe8_4d80_93f4_b9f438b6eb0e+mw9a5baf6d_0285_4ad3_9499_059c553d9cf6)
mwf0b9efb6_f0e9_4704_b5b1_dec2a68c3321=0.48 1/secondReaction: mw8bffd47e_34de_4738_81bf_7a39a40b3ae8 => mw308b75ec_28b7_4d97_92e2_51a8ce04116a; mw8bffd47e_34de_4738_81bf_7a39a40b3ae8, Rate Law: mwf0b9efb6_f0e9_4704_b5b1_dec2a68c3321*mw8bffd47e_34de_4738_81bf_7a39a40b3ae8
mw18baeb4d_ad18_4c22_95c4_2ada0f618c65=1.8 nanomole; mw234b354b_eb7b_4af6_a678_9339f6b5eb8d=0.01 nanomole/secondReaction: mw280197c8_98de_43f0_bf01_0f332a1ab689 => mw3832f277_aef2_4f1d_87af_abc2a3c1a7d5; mw280197c8_98de_43f0_bf01_0f332a1ab689, Rate Law: mw234b354b_eb7b_4af6_a678_9339f6b5eb8d*mw280197c8_98de_43f0_bf01_0f332a1ab689/(mw18baeb4d_ad18_4c22_95c4_2ada0f618c65+mw280197c8_98de_43f0_bf01_0f332a1ab689)
mwb336e12c_0e62_4fff_94c0_2771b1a19065=0.2 1/secondReaction: mw64453fc5_a275_4bba_84f0_2af249b31514 => mwda4716f1_ae00_4149_aec3_12531380425a; mw64453fc5_a275_4bba_84f0_2af249b31514, Rate Law: mwb336e12c_0e62_4fff_94c0_2771b1a19065*mw64453fc5_a275_4bba_84f0_2af249b31514

States:

NameDescription
mw3832f277 aef2 4f1d 87af abc2a3c1a7d5[Tyrosine-protein kinase JAK1]
mwbee11634 55df 4a3f 998a 634dfaf46fd7[Mitogen-activated protein kinase 8]
mw13651143 feb5 49a5 adab 9105c2647446[Signal transducer and activator of transcription 1-alpha/beta]
mw97345a67 a8e8 42aa 8e62 69e9d2b6cf45[mitogen-activated protein kinase p38 binding]
mwf20834c8 a115 460b 859c 4e3ca1ffd953[diglyceride]
mw46ee629a dd6b 4163 9da1 2614bb1d74bc[Dual specificity mitogen-activated protein kinase kinase 3; Protein kinase byr1]
mw6ee00a71 ab68 454b b1cd 60c1ebd19cfa[tumor necrosis factor receptor superfamily complex; Receptor-interacting serine/threonine-protein kinase 1; Tumor necrosis factor receptor type 1-associated DEATH domain protein; TNF receptor-associated factor 2]
mw5c67812a 17f5 43cf 8acb 9bde272c1911[Proto-oncogene c-Fos]
mw702be69a eb4f 425e 87c7 ef7d85254536[Mitogen-activated protein kinase kinase kinase 7]
mw05469f51 73f7 4ba1 9f1a bce5fea143c2[1-phosphatidyl-1D-myo-inositol 4,5-bisphosphate]
mw2dc73059 a841 48d5 b4bd 3ac24d94c42e[IkappaB kinase complex]
mw9bb804c9 3e4e 4684 9f6b 4e6f6706a58e[Phosphatidylinositol 4-phosphate 3-kinase C2 domain-containing subunit alpha]
mwe5fade7d 1715 4bb1 843f 923da8ecddf1[Myeloid differentiation primary response protein MyD88]
mw8cc67de0 64e6 428f ab09 4c2825cc172c[Tumor necrosis factor receptor superfamily member 1A]
mw262497ec 3d54 4367 bfe3 76a9c57497cb[Interleukin-1 receptor-associated kinase 1]
mw4079e13c 446e 4aa2 9ec4 233583833d02[Toll-Like Receptors Cascades; Monocyte differentiation antigen CD14; Toll-like receptor 2]
mw9a5baf6d 0285 4ad3 9499 059c553d9cf6[1-phosphatidylinositol 4,5-bisphosphate phosphodiesterase gamma-1; cytokine-mediated signaling pathway]
mw64453fc5 a275 4bba 84f0 2af249b31514[Tyrosine-protein kinase BTK]
mw7204ab72 2ee5 4b92 b420 2583dacc4343[I-kappaB phosphorylation]
mwda4716f1 ae00 4149 aec3 12531380425a[Tyrosine-protein kinase BTK]
mw4d2e70a7 f499 461d ae18 bc53b365b091[Tumor necrosis factor]
mwd9e7a9b9 6f1b 4bbc afa5 6cb192b62ce8[Mitogen-activated protein kinase 8]
mw6939cefe e7ff 4a3f b45b a9234d1b5573[Nuclear factor NF-kappa-B p105 subunit]
mw280197c8 98de 43f0 bf01 0f332a1ab689[Epidermal growth factor receptor]
mwf8cfed1b 6fcf 4cba bc30 b44490814a7a[Mitogen-activated protein kinase kinase kinase 1]
mwa5d6f7e4 dc4d 4931 91ce 1e78e7b2f195[lipophosphoglycan]
mw308b75ec 28b7 4d97 92e2 51a8ce04116a[Mitogen-activated protein kinase kinase kinase 7; TGF-beta-activated kinase 1 and MAP3K7-binding protein 3]
mw8bffd47e 34de 4738 81bf 7a39a40b3ae8[TNF receptor-associated factor 6]
mw32c21c39 237b 4d4c bb5d 117cb30ce68a[RAF proto-oncogene serine/threonine-protein kinase]
mw136c8391 14f4 4a28 83a3 35cc74a2e040[NF-kappaB-inducing kinase activity]
mw17ae9adc 54ab 407b a34d 8413a3a10cc6[Signal transducer and activator of transcription 1-alpha/beta]
mw67d0cf04 d6a7 4725 a869 098a96a3350d[Mitogen-activated protein kinase 3; Mitogen-activated protein kinase CPK1]
mw173d8585 5817 4b4c 932a cf7d673680ac[IPR020849]
mw1f12e5bc ebbc 4347 b6b7 5cd1740ac69a[Mitogen-activated protein kinase 3; Mitogen-activated protein kinase CPK1]
mw8a358487 b18b 42df a646 cd75eb5bfcc2[Nuclear factor NF-kappa-B p105 subunit]
mwb4633da9 f9d6 4ad8 a7e5 da075c830e17[Protein kinase C alpha type; Putative serine/threonine-protein kinase-like protein CCR3]
mw323a57b4 8e59 4116 9ad1 fe547b89c858[Growth factor receptor-bound protein 2; Son of sevenless homolog 2; IPR000980]
mw75377e12 e23d 44b3 9823 5fac9b23edc8[Dual specificity mitogen-activated protein kinase kinase 2; Protein kinase byr1]
mw805b55df cc91 4227 bb52 930e961b682c[Mitogen-activated protein kinase kinase kinase 5]
mwfed5a135 c91b 4d20 91b2 3a61723544dd[Transcription factor AP-1]
mwb71eb539 dca6 47ab 8df5 430d84af0bfb[mitogen-activated protein kinase p38 binding]
mw0be0d193 fd6b 4824 8928 dbade8b5c99c[Pro-epidermal growth factor]
mwc844b7c0 98f5 4d0d 8f0c 00dfe8b54e6d[Tumor necrosis factor]