Loading Controls for Western Blots
Mary Johnson (han at labome dot com)
Synatom Research, Princeton, New Jersey, United States
DOI
//dx.doi.org/10.13070/mm.en.2.114
Date
last modified : 2023-12-25; original version : 2012-02-18
Cite as
MATER METHODS 2012;2:114
Abstract

When comparing protein abundance across treatment groups using Western blotting, it is critical to have a method to account for variation due to errors in loading or protein transfer. A common approach is the use of a loading control, a constitutively expressed protein that does not vary among treatment groups, to normalize the relative expression of the protein of interest. Selection of an appropriate loading control is important and can be difficult. Here we present a comprehensive review of loading controls for Western blots, and the survey results from formal publications.

Loading Controls in Western Blots

Western blotting is commonly used to investigate the change in abundance of a specific protein under different conditions. The Western blotting process involves multiple steps, including sample preparation, sample loading, electrophoresis, protein transfer to a membrane, antibody incubation, and signal detection. To interpret the results from any Western blot experiment, a loading control is critical.

Loading Controls for Western Blots figure 1
Figure 1. Loading controls: A, beta-actin. This image is adapted from Figure 1 in [1]. B, tubulin. This image is adapted from Figure 1 in [2]. C, GAPDH. This image is adapted from Figure 2S in [3]. D, lamin B. This image is adapted from Figure 1 in [2]. E, HSC70. This image is adapted from Figure 2 in [4]. F, vinculin. This image is adapted from Figure 3 in [5].

A loading control accounts for potential variation in:

  • - the amount of protein sample loaded in each lane,
  • - protein transfer efficiencies from the gel to the membrane among different samples/lanes,
  • - antibody incubation (for primary or secondary antibody), and signal detection across separate samples/lanes.

The signals from loading controls are typically used to normalize the signals from the proteins of interest. To use a loading control for these purposes detection with control protein antibody and the experimental antibody should be done on the same blot. A variety of different proteins are used as loading controls. For any particular experiment it is important to consider potential loading controls and try to find a protein that is expressed at stable levels among the samples of interest. For this reason housekeeping genes that are constitutively expressed and necessary for basic cell functions, like cytoskeletal proteins, are typically used for loading controls. Looking to the literature can be very useful in considering potential loading controls for your Western blotting experiments.

Labome has surveyed publications with loading controls in Western blots (Table 1). Figure 1 displays the Western blot images compiled from some of those publications. The results indicate that actin (specifically beta actin) is the most commonly used control. Other interesting observations from this survey are discussed in this article later.

Supplier Num Sample Catalog Number with Reference
actin
MilliporeSigma42 A3854 A5316 [7] ; A1978 [8, 9]
Santa Cruz Biotechnology17 sc-81178 [10] ; sc-69879 / AC-15 [11]
Abcam4 ab8226 [12, 13]
Cell Signaling3 4970 [14-16]
Novus Biologicals1 NB600-501 [17]
MP Biomedicals1ICN691001 [18]
tubulin
MilliporeSigma16 05-661 [19, 20], T5326 [21]
Santa Cruz Biotechnology6 sc-53029 [22]
BioLegend2 MMS-435P [23, 24]
Abcam1 ab6160 [25]
Cell Signaling Technology1 3873 [26]
Table 1. The number of publications (num) that cites each type of main loading controls with antibodies from each of the major suppliers.

The proteins serving as loading controls must fit certain criteria. These criteria are:

  • the protein levels of loading controls remain constant (relative to the total protein content) under the test conditions (such as across treatments or developmental stages),
  • the detection bands of loading controls must not interfere with those of the protein(s) of interest (i.e., they should have substantially different molecular weights),
  • the detection limits for the loading controls and the proteins of interest are within dynamic ranges. That is, the protocol and detection method can reveal the changes in the levels of loading controls and the proteins of interest without signal saturation.

The dynamic detection range for a loading control can be determined by running a Western blot specifically for the protein control in serial dilution. The signal intensity must correlate directly with the protein concentration within the range that will be used in the experiment.

Loading Controls for Different Samples

Different sample preparations require different loading controls. Table 2 summarizes commonly used loading controls for whole cell/cytoplasmic proteins, mitochondria, nuclear proteins, plant tissues, and serum samples. Each of them is discussed in detail below.

sample typeproteinMW (kDa)
whole cell / cytoplasmic proteins
beta actin43
alpha actin43
GAPDH30-40
beta-tubulin55
alpha-tubulin55
(high molecular weight)vinculin [27] 116
mitochondria
VDCA1/porin31
cytochrome C oxidase16
nuclear proteins
lamin B166
TATA binding protein TBP38
PCNA29
histone H130
histone H318
plant tissue
LHCP25
APX332
serum
transferrin77
muscle
SDHA [28] 73
yeast
phosphoglycerate kinase45
Table 2. Commonly used loading controls for different sample preparations and their molecular weights.
Whole cell/cytoplasmic proteins
Actins

Actins are a family of six proteins in three groups in human and other vertebrates: alpha cardiac muscle 1 ( ACTC1), alpha 1 ( ACTA1, skeletal muscle) and 2 ( ACTA2, aortic smooth muscle), beta ( ACTB), gamma 1 ( ACTG1) and 2 ( ACTG2, enteric smooth muscle). Beta and gamma 1 are two non-muscle actin proteins. They function as the main components of microfilaments (Figure 2). Actins are highly-conserved. Beta-actin proteins from species as diverse as human, mouse, and chicken are identical. Human beta actin shares close to 90% identity with fungal homologs. Human beta actin is at least 93% identical with other members of the human actin family. This homology among actins and across species is important for antibody selection and the interpretation of Western blot bands.

Loading Controls for Western Blots figure 2
Figure 2. Actins and microfilaments. The image is from NIH.

Both beta and alpha actins have been used as loading controls in Western blot experiments. Please see the detailed discussion about beta-actin loading control and commonly used clones, and frequently asked questions about beta-actin controls.

Actin proteins tend to be present in cells at very high concentrations, and thus if this control is used it is essential to ensure that you are working within the dynamic detection range so there is no signal saturation and variation in actin protein levels can be detected.

Before selecting an actin protein as your loading control you should ensure that it is an appropriate control with very little to no variation among your treatment groups. Changes in cell growth conditions do disrupt actin protein synthesis. Publications have indicated that beta-actin may not be a reliable loading control in Western blot analysis in most cases [29]. Beta-actin is also present in the nucleus, as a component of chromatin remodeling complexes [30], but it can not be used as a control for nuclear protein samples.

Glyceraldehyde 3-phosphate dehydrogenase

Glyceraldehyde 3-phosphate dehydrogenase (GAPDH, G3PD, GAPD; Figure 3) catalyzes the reversible oxidative phosphorylation of glyceraldehyde-3-phosphate in the presence of inorganic phosphate and nicotinamide adenine dinucleotide (NAD) in glycolysis. It also has nitrosylase activity and is involved in gene transcription, RNA transport, DNA replication, and apoptosis.

Loading Controls for Western Blots figure 3
Figure 3. The tetramer of GAPDH is involved in glycolysis. The image is from NIH.

GAPDH is a housekeeping gene and used as controls in both Western blot and qPCR. GAPDH is highly conserved across species. For instance, human GAPDH with 335 amino acids shares about 70% identity with its 422 amino acids homolog from Arabidopsis thaliana. It is present in the cytosol, nucleus, perinuclear regions, and membranes. A comprehensive study indicates that GAPDH mRNA levels differ significantly among tissue types, but remain constant with regards to age and gender [31]. Hypoxia can upregulate GAPDH expression [32, 33], and GAPDH can aggregate [34]. Thus it may be a useful loading control in some Western blot analyses but can not be reliable in other cases, such as for oxygen-related studies. Interestingly, it is the target of dimethyl fumarate, an immunomodulatory drug used to treat psoriasis and multiple sclerosis, which succinates and inactivates the catalytic cysteine of GAPDH [35]. In addition, GAPDH is an RNA-binding protein [36].

Tubulins

There are five types of tubulin: alpha, beta, gamma, delta, and epsilon. Alpha-, beta-, and gamma-tubulins have all been used as loading controls.

Loading Controls for Western Blots figure 4
Figure 4. Tubulins as the structural components of microtubules and the targets of drugs. The image is from NIH.

Heterodimers of alpha- and beta-tubulins are the building blocks of microtubules (Figure 4). Subtypes of alpha-tubulins include 1a (TUBA1A), 1b (TUBA1B), 1c (TUBA1C), 3c (TUBA3C), 3d (TUBA3D), 3e (TUBA3E), 4a (TUBA4A), and 8 (TUBA8). The subtypes of beta-tubulins are Classes I (TUBB), IIa (TUBB2A), IIb (TUBB2B), III (TUBB3), IVa (TUBB4A), IVb (TUBB4B), V (TUBB6), VI (TUBB1), and VIII (TUBB8).

Gamma-tubulins, gamma 1(TUBG1) and 2 (TUBG2), are involved in the nucleation and polar orientation of microtubules, and are present in centrosomes and spindle pole bodies. Delta (TUBD1) and epsilon (TUBE1) tubulins tend to localize at centrioles and are structural components of mitotic spindles.

Tubulins are conserved across species. For example, human tubulin alpha 1a with 451 amino acids share 74% identity with its yeast homolog Tub3p with 445 amino acids. Subtypes of alpha, beta, or gamma are very similar as well. Human alpha-tubulins share more than 90% identity with one another. Different types of tubulins share significant homology as well. Human alpha-tubulins share 40% identity with human beta-tubulins.

Tubulins are pharmacological targets for several groups of drugs, like anti-cancer drugs Taxol (Figure 4), Tesetaxel, vinblastine, and vincristine, anti-gout agent colchicine, and anti-fungal drug griseofulvin. These drugs affect tubulin expression in vitro and in vivo.

Vinculin - high molecular weight protein loading control

Another cytoskeletal protein, vinculin, has been used as a Western blot loading control as well. It is one of the major components of cell-cell and cell-matrix junctions. It is a large protein; the molecular weight for human vinculin is 117 kDa with 1066 amino acids. It can be used as a loading control for high molecular weight proteins. In muscle tissues, a splice variant with an extra exon with the molecular weight of 150 kDa is also expressed.

Nuclear proteins - nuclear loading controls
Lamins

Lamin proteins are the structural components of the nuclear lamina (Figure 5), which support nuclear envelope and is involved in the breakdown and re-formation of the nuclear envelope during the cell cycle. In animal cells, three genes encode at least seven lamins. LMNA gene encodes both A- and C-types of lamins through alternative splicing, while LMNB1 and LMNB2 genes encode lamin B1 and B2. B-type lamins are present in every cell. A-type lamins are expressed after gastrulation, and the expression of C-type lamins is tissue-specific.

Loading Controls for Western Blots figure 5
Figure 5. Nuclear envelope and lamina. The figure is from [6].

Post-translational modifications such as farnesylation and phosphorylation occur on lamins, which in turn regulate the assembly of the nuclear lamina and the activity of lamin proteins.

Among the lamins, Lamin B1 is most commonly used as a loading control. Lamin B1 is conserved across species. Human lamin B1 shares 95% identity with rodent homologs, and 36% identity with the fruit fly counterpart. Human lamin B1 is more than 50% identical to other human lamins. However, Lamin B1 protein is not suitable as a loading control for samples without a nuclear envelope.

TATA-box binding protein

TATA-box binding protein, TBP, is a general transcription factor that binds specifically to the TATA box DNA sequence before gene transcription by RNA polymerase II. TATA boxes are present in 10-20% of human gene promoters. TBP is highly conserved. Human TBP shares about 90% identity with rodent homologs and over 60-70% identity with fungal homologs. The N-terminus of TBP contains a long string of glutamines (CAG repeats in TBP mRNA molecules), which modulates the DNA binding activity of the C-terminus. The number of N-terminal glutamines in TBP varies in healthy individuals (between 25 to 42 repeats) and expands to 47-63 repeats in certain pathological conditions. This and other post-translational modifications may cause small variation the molecular weight of TBP on Western blots.

Heat shock protein

Heat shock proteins (HSPs), also called heat shock cognates (HSCs), are a group of proteins which are upregulated under certain stressors, especially under high heat. These proteins have a wide array of functions including providing protection against such stressors and chaperoning proteins around the cell. Some HSPs, such as HSC70 and HSP90 [37], are occasionally used as Western blot loading controls. For example, Galmozzi A et al ensured equal loading of brown adipose tissue with GeneTex anti-HSP90 antibody ( GTX101423) [38]. Wang C et al used HSP90 as a loading control for Hep3B, Huh7, and other human liver cancer cell lines with different treatments [39]. While some HSPs may be constitutively expressed and thus may serve as a reliable loading control in some Western blot analyses, it is critical to keep in mind that treatment conditions may alter expression of certain HSPs.

Proliferating cell nuclear antigen

Proliferating cell nuclear antigen (PCNA) is involved in eukaryotic DNA replication. It is a cofactor of DNA polymerase delta, and is expressed in the nuclei of cells during the DNA synthesis phase (S phase) of the cell cycle. PCNA is highly conserved among chimpanzee, dog, cow, mouse, rat, chicken, zebrafish, fruit fly, mosquito, S. pombe, S. cerevisiae, K. lactis, E. gossypii, M. grisea, N. crassa, A. thaliana, and rice. Human PCNA, with 261 amino acids, shares 97% identity with rodent homologs and 36% identity with yeast homologs. However, because the function of PCNA is in DNA replication it is a poor loading control for non-proliferating cells or cells with anti-proliferation treatment.

Mitochondria
VDAC1/porin

Voltage-dependent anion channels (VDCAs), a class of porin family proteins, constitute the major proteins of the outer mitochondrial membrane in eukaryotic cells. There are at least three members (VDAC1, VDAC2, and VDAC3) in vertebrates. VDAC1 is present in both the outer mitochondrial membrane and the plasma membrane, serving different functions in each location. It is expressed in heart, liver, and muscle tissues. Multiple variants of VDAC1 exist due to alternative splicing. The VDAC1 gene is conserved among chimpanzee, dog, cow, mouse, rat, chicken, and zebrafish. Human VDAC1 share 99% identity with the mouse homolog and 85% identity with its zebrafish homolog.

Cytochrome C oxidase/COX4I1

Cytochrome C oxidase, in the inner mitochondrial membrane, is the last enzyme complex of the mitochondrial electron transport chain. The complex consists of 13 subunits encoded by both mitochondrial and nuclear genes.

Subunit IV of cytochrome C oxidase has two isoforms (isoforms 1 and 2). They are encoded by two nuclear genes (COX4I1 and COX4I2). Isoform 1 is ubiquitously expressed, and isoform 2 is highly expressed in lung tissues. Isoform I COX4I1 is frequently used as a loading control. Human COX4I1 shares 80% identity with the mouse homolog, 63% identical to its zebrafish homolog, and 60% identity with human COX4I2.

Extracellular fluids, urine, blood, cell culture media, sperms
Transferrin for serum samples

Transferrin is a blood plasma glycoprotein sometimes used as a Western blot loading control. Its major function is to transport iron from the liver, intestines and reticuloendothelial system to proliferating cells throughout the body. Human transferrin, with 698 amino acids, is 73% identical to its mouse homolog and 43% identical to the zebrafish homolog. When considering Transferrin as a loading control it is important to be aware that its expression is affected in some inherited diseases and by retinoic acid treatment.

Plant membranes
APX3

APX3, ascorbate peroxidase 3, has been used as a Western blot loading control for plant membrane fractions. It is a microsomal ascorbate peroxidase, which scavenges hydrogen peroxide in plant cells. In Arabidopsis there are three cytosolic (APX1, APX2, APX6), two chloroplastic types (stromal sAPX, thylakoid tAPX), and three microsomal (APX3, APX4, APX5) isoforms.

Others

Ling Q et al used Slp1 and Tic110 as loading controls for western blot in Arabidopsis protoplast lysates [40].

Other Considerations

Several authors have proposed to forego the use of a protein as a loading control, and to rely on the dye staining of proteins before (by Coomassie blue [41] ) or after (by Ponceau [42] or, more recently, REVERT [43, 44] ) the transfer step during Western blotting, or use a Stain-Free technology (adding a 58-Da Trihalo compound to the gel) [45]. The results for the protein of interest can then be compared to the total protein rather than a selected loading control that may vary unexpectedly among treatment groups, leading to inaccurate results.

The problem with such alternatives to loading controls is that they may not be able to account for all three considerations of an ideal control: sample loading, protein transfer, and antibody incubation/signal detection. Therefore, a combination of these processes, a protein loading control, Coomassie blue staining of the gel, and Ponceau staining of the transfer membrane, constitutes a well-devised Western blot protocol and thus should be considered. In this way you can confirm your observed results are consistent and thus more robust.

Antibodies for Typical Loading Controls in the Literature

Antibodies from many suppliers have been used to detect different types of loading controls among the publications Labome surveyed.

Actin antibodies

Santa Cruz anti-beta-actin antibodies served as loading controls in A549 whole cell lysates, HUVEC and dermal fibroblast lysates ( sc-47778) [21, 46] and intestinal epithelial cells from mouse terminal ileum [47].

MilliporeSigma anti-actin antibodies were used for loading controls in Western blots examination [48, 49].

Abcam anti-actin antibodies were used in studies involving K562-TP53 isogenic AML cell lines [13], mouse peritoneal macrophage and human RAW264.7 cells [12].

Thermo Fisher anti-beta-actin antibody was used as a loading control with human carcinoma cell lines to investigate the role of Brachyury in tumor cell epithelial-mesenchymal transition and tumor progression [50].

Chakraborty AA et al used Cell Signaling actin antibody (catalog: 4970) at 1:5000 as a loading control with human mammary epithetelial cells [15]. Zeng Q et al used the same antibody as a loading control for the western blot analysis of N-methyl-D-aspartate receptors in human MDA231 and mouse TS1 breast cancer cells and their corresponding breast-to-brain metastasis derivative cells [14].

Tubulins

In the articles examined by Labome, alpha-, beta-, and gamma-tubulins have been used as loading controls.

Santa Cruz Biotechnology anti-alpha-tubulin antibody was used in an optoSOS study (catalog number: sc-23948) [51] and in transfected U937 cells [52]. Patzke C et al used immunoblotting with anti-tubulin beta III mouse monoclonal TuJ1 antibody from BioLegend / Covance as a loading control for human induced neuron lysates [24].

MilliporeSigma anti-tubulin antibodies were used as loading controls with NIH3T3 cells 05-661 [20], A549 cells (a pd:4409345>T5326) [21], HEK293T [53], and C2C12 cells [15]. Hamilton WB et al detected tubulin in ES cell lysates with Abcam ab6160 to serve as a loading control [25].

GAPDH

Moya IM et al used antibody AM4300 from Thermo Fisher to stain GAPDH in Western blot to ensure equal loading of mouse liver lysates [54]. Laflamme C et al used OriGene antibody ( TA802519) to stain the lysates from HEK-293 cells [44]. Santa Cruz Biotechnology anti-GAPDH antibody was used as loading controls in HepG2 and HeLa cell lysates [16] or lysates of primary myoblasts and fibroblasts ( sc-47724) [55]. Abcam monoclonal anti-GAPDH antibody ( ab8245) was used in Western blotting as a loading control in choroid plexus organoids [56] and 3T3 cell lysate [57]. Li L et al used the GAPDH antibody from Yeasen (

30201ES20) [58]. MilliporeSigma anti-GAPDH antibodies were used as controls in lysates of U2OS cell lines [59], zebrafish ventricles [60], 293T cells expressing pseudotyped viruses ( G8795) [61], human circulating tumor cells ( ABS16 [62] ), ES cells ( G8795) [25], extracted mouse epithelial cells [63] and in the cytoplasmic fractions of mouse livers and HepG2 cells ( MAB374) [49].

GAPDH antibodies from CST clone 14C10 [46, 64, 65], D46CR [66] and D16H11 [67] were used as well. Choi JH et al used D16H11 GAPDH antibodies from CST as a cytosolic marker [67]. Zeng Q et al used the CST GAPDH antibody ( 2118) as a loading control for the western blot analysis of N-methyl-D-aspartate receptors in human and mouse breast cancer cells and their corresponding breast-to-brain metastasis derivative cells [14].

Vinculin

de Morree A et al compared the expression of Pax3 protein in muscle stem cells between wild-type and Pax3-KO mice and among the wild-type mice treated with various antisense vivo-morpholino oligonucleotides in Peggy Sue capillary western using vinculin detected by MilliporeSigma V9131) as a loading control [68]. Lundby A et al detected A549 whole cell lysates with anti-vinculin antibody V9264 from MilliporeSigma [21]. Chakraborty AA et al used vinculin antibody from MilliporeSigma ( V9131) at 1:5000 on C2C12 cell lysates [15]. Vasan N et al measured the sample loading of MCF10A, NIH-3T3 and MCF7 cell lysates with the anti-vinculin antibody (13901) from from Cell Signaling Technology [69].

Others

A variety of other protein targets have been selected to serve as loading controls under a myriad of circumstances. Li Z et al used spectrin, detected with the MilliporeSigma MAB1622 antibody as a control for the lysates from mouse cortices and cultured mouse cortical neurons [70]. Zeitler B et al used calnexin as a loading for lysates of Huntington’s disease fibroblast line GM04723 cells [71]. Engle DD et al normalized Western blot signals from organoid lysates against cofilin, an actin-binding protein which disassembles actin filaments [72]. Tornabene P et al measured protein expression in retinal lysates from mice and pigs with dysferlin detected with clone Ham1/7B6 (MONX10795) from Tebu-bio as controls [17]. Choi JH et al used histone H3 with Cell Signaling Technology antibody ( 9715) as a nuclear fraction marker for Western blot [67]. Saito T et al used lamin B as a nuclear fraction marker for lysates from mouse livers and HepG2 cells [49]. Transferrin was used as a loading control for serum samples [73]. The papers examined by Labome also found lamin A/C [74], histone H3 [67, 75] used as Western blot loading controls.

References
  1. Huang D, Wang Y, Wang L, Zhang F, Deng S, Wang R, et al. Poly(ADP-ribose) polymerase 1 is indispensable for transforming growth factor-? Induced Smad3 activation in vascular smooth muscle cell. PLoS ONE. 2011;6:e27123 pubmed publisher
  2. Mendoza Villanueva D, Deng W, Lopez Camacho C, Shore P. The Runx transcriptional co-activator, CBFbeta, is essential for invasion of breast cancer cells. Mol Cancer. 2010;9:171 pubmed publisher
  3. Hay E, Nouraud A, Marie P. N-cadherin negatively regulates osteoblast proliferation and survival by antagonizing Wnt, ERK and PI3K/Akt signalling. PLoS ONE. 2009;4:e8284 pubmed publisher
  4. Pereira E, Liao N, Neale G, Hendershot L. Transcriptional and post-transcriptional regulation of proangiogenic factors by the unfolded protein response. PLoS ONE. 2010;5: pubmed publisher
  5. Schiappacassi M, Lovisa S, Lovat F, Fabris L, Colombatti A, Belletti B, et al. Role of T198 modification in the regulation of p27(Kip1) protein stability and function. PLoS ONE. 2011;6:e17673 pubmed publisher
  6. Chi Y, Chen Z, Jeang K. The nuclear envelopathies and human diseases. J Biomed Sci. 2009;16:96 pubmed publisher
  7. Orthofer M, Valsesia A, Magi R, Wang Q, Kaczanowska J, Kozieradzki I, et al. Identification of ALK in Thinness. Cell. 2020;181:1246-1262.e22 pubmed publisher
  8. Daly J, Simonetti B, Klein K, Chen K, Williamson M, Antón Plágaro C, et al. Neuropilin-1 is a host factor for SARS-CoV-2 infection. Science. 2020;370:861-865 pubmed publisher
  9. Genet G, Boyé K, Mathivet T, Ola R, Zhang F, Dubrac A, et al. Endophilin-A2 dependent VEGFR2 endocytosis promotes sprouting angiogenesis. Nat Commun. 2019;10:2350 pubmed publisher
  10. Zhang T, Xu D, Trefts E, Lv M, Inuzuka H, Song G, et al. Metabolic orchestration of cell death by AMPK-mediated phosphorylation of RIPK1. Science. 2023;380:1372-1380 pubmed publisher
  11. Johmura Y, Yamanaka T, Omori S, Wang T, Sugiura Y, Matsumoto M, et al. Senolysis by glutaminolysis inhibition ameliorates various age-associated disorders. Science. 2021;371:265-270 pubmed publisher
  12. Wang L, Wen M, Cao X. Nuclear hnRNPA2B1 initiates and amplifies the innate immune response to DNA viruses. Science. 2019;365: pubmed publisher
  13. Boettcher S, Miller P, Sharma R, McConkey M, Leventhal M, Krivtsov A, et al. A dominant-negative effect drives selection of TP53 missense mutations in myeloid malignancies. Science. 2019;365:599-604 pubmed publisher
  14. Zeng Q, Michael I, Zhang P, Saghafinia S, Knott G, Jiao W, et al. Synaptic proximity enables NMDAR signalling to promote brain metastasis. Nature. 2019;573:526-531 pubmed publisher
  15. Chakraborty A, Laukka T, Myllykoski M, Ringel A, Booker M, Tolstorukov M, et al. Histone demethylase KDM6A directly senses oxygen to control chromatin and cell fate. Science. 2019;363:1217-1222 pubmed publisher
  16. Huang H, Weng H, Zhou K, Wu T, Zhao B, Sun M, et al. Histone H3 trimethylation at lysine 36 guides m6A RNA modification co-transcriptionally. Nature. 2019;567:414-419 pubmed publisher
  17. Tornabene P, Trapani I, Minopoli R, Centrulo M, Lupo M, de Simone S, et al. Intein-mediated protein trans-splicing expands adeno-associated virus transfer capacity in the retina. Sci Transl Med. 2019;11: pubmed publisher
  18. Mena E, Kjolby R, Saxton R, Werner A, Lew B, Boyle J, et al. Dimerization quality control ensures neuronal development and survival. Science. 2018;362: pubmed publisher
  19. Lu Y, Brommer B, Tian X, Krishnan A, Meer M, Wang C, et al. Reprogramming to recover youthful epigenetic information and restore vision. Nature. 2020;588:124-129 pubmed publisher
  20. Schüchner S, Behm C, Mudrak I, Ogris E. The Myc tag monoclonal antibody 9E10 displays highly variable epitope recognition dependent on neighboring sequence context. Sci Signal. 2020;13: pubmed publisher
  21. Lundby A, Franciosa G, Emdal K, Refsgaard J, Gnosa S, Bekker Jensen D, et al. Oncogenic Mutations Rewire Signaling Pathways by Switching Protein Recruitment to Phosphotyrosine Sites. Cell. 2019;179:543-560.e26 pubmed publisher
  22. Pataskar A, Champagne J, Nagel R, Kenski J, Laos M, Michaux J, et al. Tryptophan depletion results in tryptophan-to-phenylalanine substitutants. Nature. 2022;: pubmed publisher
  23. Cole T, Zhao H, Collier T, Sandoval I, Sortwell C, Steece Collier K, et al. α-Synuclein antisense oligonucleotides as a disease-modifying therapy for Parkinson's disease. JCI Insight. 2021;6: pubmed publisher
  24. Patzke C, Brockmann M, Dai J, Gan K, Grauel M, Fenske P, et al. Neuromodulator Signaling Bidirectionally Controls Vesicle Numbers in Human Synapses. Cell. 2019;179:498-513.e22 pubmed publisher
  25. Hamilton W, Mosesson Y, Monteiro R, Emdal K, Knudsen T, Francavilla C, et al. Dynamic lineage priming is driven via direct enhancer regulation by ERK. Nature. 2019;: pubmed publisher
  26. Guiley K, Stevenson J, Lou K, Barkovich K, Kumarasamy V, Wijeratne T, et al. p27 allosterically activates cyclin-dependent kinase 4 and antagonizes palbociclib inhibition. Science. 2019;366: pubmed publisher
  27. Barnat M, Capizzi M, Aparicio E, Boluda S, Wennagel D, Kacher R, et al. Huntington's disease alters human neurodevelopment. Science. 2020;369:787-793 pubmed publisher
  28. Wai T, García Prieto J, Baker M, Merkwirth C, Benit P, Rustin P, et al. Imbalanced OPA1 processing and mitochondrial fragmentation cause heart failure in mice. Science. 2015;350:aad0116 pubmed publisher
  29. Dittmer A, Dittmer J. Beta-actin is not a reliable loading control in Western blot analysis. Electrophoresis. 2006;27:2844-5 pubmed
  30. Olave I, Reck Peterson S, Crabtree G. Nuclear actin and actin-related proteins in chromatin remodeling. Annu Rev Biochem. 2002;71:755-81 pubmed
  31. Barber R, Harmer D, Coleman R, Clark B. GAPDH as a housekeeping gene: analysis of GAPDH mRNA expression in a panel of 72 human tissues. Physiol Genomics. 2005;21:389-95 pubmed
  32. Yamaji R, Fujita K, Takahashi S, Yoneda H, Nagao K, Masuda W, et al. Hypoxia up-regulates glyceraldehyde-3-phosphate dehydrogenase in mouse brain capillary endothelial cells: involvement of Na+/Ca2+ exchanger. Biochim Biophys Acta. 2003;1593:269-76 pubmed
  33. Zhong H, Simons J. Direct comparison of GAPDH, beta-actin, cyclophilin, and 28S rRNA as internal standards for quantifying RNA levels under hypoxia. Biochem Biophys Res Commun. 1999;259:523-6 pubmed
  34. Nakajima H, Itakura M, Kubo T, Kaneshige A, Harada N, Izawa T, et al. Glyceraldehyde-3-phosphate Dehydrogenase (GAPDH) Aggregation Causes Mitochondrial Dysfunction during Oxidative Stress-induced Cell Death. J Biol Chem. 2017;292:4727-4742 pubmed publisher
  35. Kornberg M, Bhargava P, Kim P, Putluri V, Snowman A, Putluri N, et al. Dimethyl fumarate targets GAPDH and aerobic glycolysis to modulate immunity. Science. 2018;360:449-453 pubmed publisher
  36. Garcin E. GAPDH as a model non-canonical AU-rich RNA binding protein. Semin Cell Dev Biol. 2019;86:162-173 pubmed publisher
  37. Kim H, Kim J, Yu S, Lee Y, Park J, Choi R, et al. A Mechanism for microRNA Arm Switching Regulated by Uridylation. Mol Cell. 2020;78:1224-1236.e5 pubmed publisher
  38. Galmozzi A, Kok B, Kim A, Montenegro Burke J, Lee J, Spreafico R, et al. PGRMC2 is an intracellular haem chaperone critical for adipocyte function. Nature. 2019;576:138-142 pubmed publisher
  39. Wang C, Vegna S, Jin H, Benedict B, Lieftink C, Ramirez C, et al. Inducing and exploiting vulnerabilities for the treatment of liver cancer. Nature. 2019;: pubmed publisher
  40. Ling Q, Broad W, Trösch R, Töpel M, Demiral Sert T, Lymperopoulos P, et al. Ubiquitin-dependent chloroplast-associated protein degradation in plants. Science. 2019;363: pubmed publisher
  41. Welinder C, Ekblad L. Coomassie staining as loading control in Western blot analysis. J Proteome Res. 2011;10:1416-9 pubmed publisher
  42. Romero Calvo I, Ocón B, Martínez Moya P, Suárez M, Zarzuelo A, Martinez Augustin O, et al. Reversible Ponceau staining as a loading control alternative to actin in Western blots. Anal Biochem. 2010;401:318-20 pubmed publisher
  43. Biever A, Glock C, Tushev G, Ciirdaeva E, Dalmay T, Langer J, et al. Monosomes actively translate synaptic mRNAs in neuronal processes. Science. 2020;367: pubmed publisher
  44. Laflamme C, McKeever P, Kumar R, Schwartz J, Kolahdouzan M, Chen C, et al. Implementation of an antibody characterization procedure and application to the major ALS/FTD disease gene C9ORF72. elife. 2019;8: pubmed publisher
  45. Vigelsø A, Dybboe R, Hansen C, Dela F, Helge J, Guadalupe Grau A. GAPDH and β-actin protein decreases with aging, making Stain-Free technology a superior loading control in Western blotting of human skeletal muscle. J Appl Physiol (1985). 2015;118:386-94 pubmed publisher
  46. Moro A, Driscoll T, Boraas L, Armero W, Kasper D, Baeyens N, et al. MicroRNA-dependent regulation of biomechanical genes establishes tissue stiffness homeostasis. Nat Cell Biol. 2019;21:348-358 pubmed publisher
  47. Ladinsky M, Araujo L, Zhang X, Veltri J, Galán Díez M, Soualhi S, et al. Endocytosis of commensal antigens by intestinal epithelial cells regulates mucosal T cell homeostasis. Science. 2019;363: pubmed publisher
  48. Herb M, Gluschko A, Wiegmann K, Farid A, Wolf A, Utermöhlen O, et al. Mitochondrial reactive oxygen species enable proinflammatory signaling through disulfide linkage of NEMO. Sci Signal. 2019;12: pubmed publisher
  49. Saito T, Kuma A, Sugiura Y, Ichimura Y, Obata M, Kitamura H, et al. Autophagy regulates lipid metabolism through selective turnover of NCoR1. Nat Commun. 2019;10:1567 pubmed publisher
  50. Fernando R, Litzinger M, Trono P, Hamilton D, Schlom J, Palena C. The T-box transcription factor Brachyury promotes epithelial-mesenchymal transition in human tumor cells. J Clin Invest. 2010;120:533-44 pubmed publisher
  51. Bugaj L, Sabnis A, Mitchell A, Garbarino J, Toettcher J, Bivona T, et al. Cancer mutations and targeted drugs can disrupt dynamic signal encoding by the Ras-Erk pathway. Science. 2018;361: pubmed publisher
  52. Yao P, Potdar A, Ray P, Eswarappa S, Flagg A, Willard B, et al. The HILDA complex coordinates a conditional switch in the 3'-untranslated region of the VEGFA mRNA. PLoS Biol. 2013;11:e1001635 pubmed publisher
  53. Frottin F, Schueder F, Tiwary S, Gupta R, Korner R, Schlichthaerle T, et al. The nucleolus functions as a phase-separated protein quality control compartment. Science. 2019;365:342-347 pubmed publisher
  54. Moya I, Castaldo S, Van den Mooter L, Soheily S, Sansores Garcia L, Jacobs J, et al. Peritumoral activation of the Hippo pathway effectors YAP and TAZ suppresses liver cancer in mice. Science. 2019;366:1029-1034 pubmed publisher
  55. Meertens L, Hafirassou M, Couderc T, Bonnet Madin L, Kril V, Kummerer B, et al. FHL1 is a major host factor for chikungunya virus infection. Nature. 2019;: pubmed publisher
  56. Pellegrini L, Bonfio C, Chadwick J, Begum F, Skehel M, Lancaster M. Human CNS barrier-forming organoids with cerebrospinal fluid production. Science. 2020;: pubmed publisher
  57. Kempf A, Tews B, Arzt M, Weinmann O, Obermair F, Pernet V, et al. The sphingolipid receptor S1PR2 is a receptor for Nogo-a repressing synaptic plasticity. PLoS Biol. 2014;12:e1001763 pubmed publisher
  58. Li L, Lai F, Hu X, Liu B, Lu X, Lin Z, et al. Multifaceted SOX2-chromatin interaction underpins pluripotency progression in early embryos. Science. 2023;382:eadi5516 pubmed publisher
  59. Yu H, Lu S, Gasior K, Singh D, Vazquez Sanchez S, Tapia O, et al. HSP70 chaperones RNA-free TDP-43 into anisotropic intranuclear liquid spherical shells. Science. 2021;371: pubmed publisher
  60. Chávez M, Morales R, Lopez Crisosto C, Roa J, Allende M, Lavandero S. Autophagy Activation in Zebrafish Heart Regeneration. Sci Rep. 2020;10:2191 pubmed publisher
  61. Letko M, Marzi A, Munster V. Functional assessment of cell entry and receptor usage for SARS-CoV-2 and other lineage B betacoronaviruses. Nat Microbiol. 2020;: pubmed publisher
  62. Ebright R, Lee S, Wittner B, Niederhoffer K, Nicholson B, Bardia A, et al. Deregulation of ribosomal protein expression and translation promotes breast cancer metastasis. Science. 2020;: pubmed publisher
  63. Krndija D, el Marjou F, Guirao B, Richon S, Leroy O, Bellaiche Y, et al. Active cell migration is critical for steady-state epithelial turnover in the gut. Science. 2019;365:705-710 pubmed publisher
  64. Zhu P, Khatiwada S, Cui Y, Reineke L, Dooling S, Kim J, et al. Activation of the ISR mediates the behavioral and neurophysiological abnormalities in Down syndrome. Science. 2019;366:843-849 pubmed publisher
  65. Guo A, Wang Y, Chen B, Wang Y, Yuan J, Zhang L, et al. E-C coupling structural protein junctophilin-2 encodes a stress-adaptive transcription regulator. Science. 2018;362: pubmed publisher
  66. Lynn R, Weber E, Sotillo E, Gennert D, Xu P, Good Z, et al. c-Jun overexpression in CAR T cells induces exhaustion resistance. Nature. 2019;576:293-300 pubmed publisher
  67. Choi J, Zhong X, McAlpine W, Liao T, Zhang D, Fang B, et al. LMBR1L regulates lymphopoiesis through Wnt/β-catenin signaling. Science. 2019;364: pubmed publisher
  68. de Morrée A, Klein J, Gan Q, Farup J, Urtasun A, Kanugovi A, et al. Alternative polyadenylation of Pax3 controls muscle stem cell fate and muscle function. Science. 2019;366:734-738 pubmed publisher
  69. VASAN N, Razavi P, Johnson J, Shao H, Shah H, Antoine A, et al. Double PIK3CA mutations in cis increase oncogenicity and sensitivity to PI3Kα inhibitors. Science. 2019;366:714-723 pubmed publisher
  70. Li Z, Wang C, Wang Z, Zhu C, Li J, Sha T, et al. Allele-selective lowering of mutant HTT protein by HTT-LC3 linker compounds. Nature. 2019;: pubmed publisher
  71. Zeitler B, Froelich S, Marlen K, Shivak D, Yu Q, Li D, et al. Allele-selective transcriptional repression of mutant HTT for the treatment of Huntington's disease. Nat Med. 2019;25:1131-1142 pubmed publisher
  72. Engle D, Tiriac H, Rivera K, Pommier A, Whalen S, Oni T, et al. The glycan CA19-9 promotes pancreatitis and pancreatic cancer in mice. Science. 2019;364:1156-1162 pubmed publisher
  73. Minagawa I, Fukuda M, Ishige H, Kohriki H, Shibata M, Park E, et al. Relaxin-like factor (RLF)/insulin-like peptide 3 (INSL3) is secreted from testicular Leydig cells as a monomeric protein comprising three domains B-C-A with full biological activity in boars. Biochem J. 2012;441:265-73 pubmed publisher
  74. Mahadevan K, Zhang H, Akef A, Cui X, Gueroussov S, Cenik C, et al. RanBP2/Nup358 potentiates the translation of a subset of mRNAs encoding secretory proteins. PLoS Biol. 2013;11:e1001545 pubmed publisher
  75. Pillow T, Adhikari P, Blake R, Chen J, Del Rosario G, Deshmukh G, et al. Antibody Conjugation of a Chimeric BET Degrader Enables in vivo Activity. ChemMedChem. 2020;15:17-25 pubmed publisher
ISSN : 2329-5139