Proteasomal and lysosomal degradation for specific and durable suppression of immunotherapeutic targets ======================================================================================================= * Yungang Wang * Shouyan Deng * Jie Xu ## Abstract Cancer immunotherapy harness the body’s immune system to eliminate cancer, by using a broad panel of soluble and membrane proteins as therapeutic targets. Immunosuppression signaling mediated by ligand-receptor interaction may be blocked by monoclonal antibodies, but because of repopulation of the membrane *via* intracellular organelles, targets must be eliminated in whole cells. Targeted protein degradation, as exemplified in proteolysis targeting chimera (PROTAC) studies, is a promising strategy for selective inhibition of target proteins. The recently reported use of lysosomal targeting molecules to eliminate immune checkpoint proteins has paved the way for targeted degradation of membrane proteins as crucial anti-cancer targets. Further studies on these molecules’ modes of action, target-binding “warheads”, lysosomal sorting signals, and linker design should facilitate their rational design. Modifications and derivatives may improve their cell-penetrating ability and the *in vivo* stability of these pro-drugs. These studies suggest the promise of alternative strategies for cancer immunotherapy, with the aim of achieving more potent and durable suppression of tumor growth. Here, the successes and limitations of antibody inhibitors in cancer immunotherapy, as well as research progress on PROTAC- and lysosomal-dependent degradation of target proteins, are reviewed. * Cancer immunotherapy * membrane protein * PROTAC * targeted degradation ## Introduction Immune checkpoint blockade therapy has shown promise in restoring the antitumor immune response through activating the immune system1,2. Multiple immunocheckpoint pathways have been described, such as those involving cytotoxic T-lymphocyte-associated protein 4 (CTLA-4), V domain-containing Ig suppressor of T-cell activation (VISTA), CD47, OX40, and T cell immunoglobulin and mucin domain 3 (TIM-3)3–5. The programed cell death 1 (PD-1)/programed cell death ligand 1 (PD-L1) axis is the most important pathway in terms of clinical therapeutic effects. PD-L1, a type I transmembrane protein containing a short cytoplasmic domain and 2 extracellular domains, is an important immune checkpoint molecule that is commonly expressed on the surfaces of cancer cells6 and is encoded by the Pdcd1 gene. The promoter region of Pdcd1 has 2 transcription-factor binding sites (termed conserved regions B and C), which are critical for regulating PD-1 expression. Pdcd1 regulation occurs partly *via* the recruitment of nuclear factor of activated T cells 1 (NFATc1) to a novel regulatory element at the Pdcd1 locus, and it is part of the molecular mechanism underlying the induction of PD-1 in response to T cell stimulation. The interaction between PD-1 on T cells and PD-L1 on cancer cells results in the suppression of tumor-killing activity of T cells, and is a crucial mechanism of tumor immune escape7. In addition, tumor-intrinsic oncogenic PD-1 promotes tumor cell proliferation independently of adaptive immunity8. Checkpoint inhibitors promote the anti-tumor immune response by antagonizing suppressive immune checkpoint regulatory pathways. PD-1/PD-L1 blockade therapy restores the effective activation of T-cell-mediated antitumor immunity and has substantial clinical benefits9. The production of antibodies targeting PD-1 and PD-L1 has led to the development of drugs targeting these pathways10. Multiple anti-PD-1/PD-L1 monoclonal antibodies (mAbs) have been approved as a therapy to treat many types of cancer11. However, PD-L1 expression varies significantly among tumor types and stages, and its level changes during therapy12. Moreover, the relatively low response rate, acquired resistance, and occasional fatal adverse effects also pose substantial challenges to use of this therapy13. The molecular mechanism regulating PD-1/PD-L1 remains largely unknown14. Small molecule-mediated inhibition of protein function is a canonical paradigm that enhances the efficacy of most clinical agents15,16. However, most small molecule inhibitors of the PD-1/PD-L1 pathway are not yet ready for widespread clinical use, and further preclinical work is required to optimize their formulation and application. Moreover, these small molecule inhibitors can achieve substantial inhibitory effects only when more than 90% of targets are engaged17. The high dosing level required can lead to on-target off-tumor effects. Gene knockdown methods based on RNA interference and CRISPR/Cas9 or related strategies have been applied to decrease cellular protein levels and have shown clear therapeutic potential18,19. Nevertheless, attaining sufficient concentrations of related agents at the target site is difficult. Safety concerns also arise from off-target effects. Moreover, these agents’ poor metabolic stability causes many adverse effects20. Therefore, PD-1/PD-L1 blockade therapy still faces great challenges that must be overcome. In-depth understanding of the control of PD-L1 expression in tumor cells is required for further improvement of checkpoint blockade therapy. PROTAC, a bifunctional small molecule compound, has been widely studied in the field of anti-cancer drugs and is used as a new therapeutic method21,22. The recently reported lysosomal targeting molecules for eliminating immune checkpoint proteins have provided a new direction for targeted degradation of target proteins as crucial anti-cancer targets23,24. Here, we primarily review the progress that has been made in degradation strategies for immunotherapeutic targets. These strategies have shown promise in providing an alternative strategy for cancer immunotherapy to achieve more potent and durable suppression of tumor growth. ## Blockade of PD-1/PD-L1 function in cancer immunotherapy ### Monoclonal antibodies Monoclonal antibodies are a promising strategy for the blockade of PD-1/PD-L1 function. Humanized mAb targeting PD-1/PD-L1 relieves T cell immunosuppression and induces T cell activation, thus restoring the body’s ability to monitor and attack tumor cells. Atezolizumab is the first licensed anti-PD-L1 mAb. Atezolizumab is designed to target PD-L1 through binding to the front beta-sheet of PD-L125,26. Atezolizumab restores the anti-tumor activity of T cells by inhibiting the interaction of PD-L1 with PD-1 on the surfaces of T cells27. Pembrolizumab, another humanized anti-PD-L1 mAb, has low affinity for Fc receptors and C1q, and a low likelihood of host immunity stimulation28. Pembrolizumab has shown strong anti-tumor activity in phase I clinical trials and is widely used in patients with advanced malignant tumors29–31. Nivolumab, a humanized anti-PD-1 IgG4 mAb, binds an N-terminal loop outside the IgV domain of PD-132. Nivolumab has been approved for application in combination with platinum-based chemotherapy33,34. With the extensive development of clinical treatments, new problems have arisen in the practical use of mAb preparations. Monoclonal antibodies induce the production of anti-mAbs, thus leading to immune-related adverse events (irAEs), such as interstitial pneumonitis, colitis with gastrointestinal perforation, and severe skin reactions35,36. In fact, PD-L1 is transported from the plasma membrane into the cytosol and actively redistributed to the plasma membrane, thereby decreasing mAb efficacy, although mAbs can effectively block PD-L1 on the surfaces of tumor cells8,37 (**Figure 1**). Emerging evidence indicates that exosomal PD-L1 mediates resistance to immunotherapy by facilitating PD-L1 evasion of the anti-PD1/PD-L1 mAbs38. Tumor-derived exosomes carry bioactive PD-L1 molecules on their surfaces that suppress the anti-tumor immune response. This exosomal PD-L1 can enter the blood circulation and inhibit T cells outside the primary tumor tissue, thus causing T cells to lose their anti-tumor ability before reaching the tumor39,40. Studies have also revealed the functions of PD-1 and PD-L1 independently of immunosuppression. The intracellular domain of PD-L1 regulates the malignant behaviors of cancer cells and mediates chemoresistance7. PD-1 is expressed in a broad range of tumor cells. The cancer-intrinsic PD-1 promotes malignant proliferation by upregulating mammalian target of rapamycin (mTOR) signaling8. These functions may explain the unexpected effects of mAbs. ![Figure 1](http://www.cancerbiomed.org/https://www.cancerbiomed.org/content/cbm/17/3/583/F1.medium.gif) [Figure 1](http://www.cancerbiomed.org/content/17/3/583/F1) Figure 1 Subcellular transportation of PD-L1 and actions of anti-PD-L1 antibody. The antibody drug binds PD-L1 expressed on the tumor cell surface, thereby blocking its interaction with PD-1. PD-L1 is degraded in the lysosome, in a process relying on several subcellular transport steps from the cell membrane to the endosome, and finally to the lysosome. PD-L1 can also be transported to recycling endosomes, thus decreasing the distribution to late endosomes and lysosomes. PD-L1 is delivered to late endosomes and then sorted to lysosomes *via* multivesicular bodies (MVBs) for degradation. ### Small molecule peptide inhibitors The disadvantages of mAbs limit their application. Although much progress has been made in the development of antibodies against the PD-1/PD-L1 pathway, there is an increasing desire to use small molecules to block the PD-1/PD-L1 axis through dissociating the PD-1/PD-L1 complex. The benefits of using small molecules instead of antibodies include better oral bioavailability, fewer irAEs, better tumor permeability, and lower production costs. The long half-life is a major limitation of mAbs. Small molecule drugs are based primarily on using different therapeutic methods to target the PD-1/PD-L1 pathway. Small molecule inhibitors are more suitable for oral administration and can decrease the target occupancy time by regulating the half-life of the drug, thereby avoiding serious irAEs41,42. Tripeptidyl peptidase 1 (TPP1), an active small-molecule peptide, has high affinity for human PD-L143. In a mouse model, TPP-1 has been found to reactivate T cells through blocking the PD-1/PD-L1 interaction and to inhibit tumor growth44. Nonylphenol ethoxylate (NP-12), a polypeptide antagonist of the PD-1 signaling pathway, is used as an immunomodulator for cancer treatment44. In mouse models of colon cancer and melanoma, NP-12 inhibits PD-1/PD-L1 interaction and suppresses tumor growth and metastasis45. CA-170, a small molecule that has been tested in clinical trials, inhibits both the PD-L1 pathway and the VISTA pathway46. Sulfamethazine and sulfamethoxazole are small molecules originally produced to inhibit the PD-1/PD-L pathway, and they have been found to rescue PD-1-mediated suppression of IFN-γ secretion47. Several other small molecule compounds that inhibit the PD-L1 pathway have been patented48. Notably, peptide inhibitors remain in an early stage of development, although they are promising in suppressing immune checkpoints. Furthermore, high drug doses are generally required, thus often leading to undesired adverse effects because of the off-target binding associated with higher drug concentrations44,45,49. ## Targeted protein degradation as a promising strategy for drug development ### Small molecule protein proteolysis-targeting chimeras as antitumor agents Small molecule inhibitors have been used to control cellular protein levels through occupancy-driven pharmacology as the mode of action. However, this strategy only temporarily inhibits the functions of regulatory proteins50,51. Finding new models is essential to control cellular protein levels. PROTACs, an attractive new approach for removing proteins by using cellular protein degradation systems to hijack the ubiquitin proteasome system, are playing an increasingly important role in drug discovery52,53. PROTACs are activators of ubiquitin ligase whose catalytic properties can be programmed (**Figure 2**)54. The heterobifunctional molecules of PROTACs recruit specific target proteins to E3 ubiquitin ligase and then reprogram the enzyme to ubiquitinate the selected target proteins, thus leading to target ubiquitination and degradation. PROTACs also activate ubiquitin ligase through mediating the formation of target protein-PROTAC-E3 ligase catalytic ternary complexes, thus providing a framework for more robust PROTAC designs55,56. ![Figure 2](http://www.cancerbiomed.org/https://www.cancerbiomed.org/content/cbm/17/3/583/F2.medium.gif) [Figure 2](http://www.cancerbiomed.org/content/17/3/583/F2) Figure 2 The modes of action of PROTACs and lysosomal targeting molecules. (A) PROTACs can bind target proteins and E3 ligase, thus forming a target protein-PROTAC-E3 ligase ternary complex, which places the target protein and E3 ligase in proximity. Ubiquitin is then transferred from E3 ligase to the target protein. Finally, the target protein is completely degraded through the action of protease. (B) The lysosomal targeting construct contains 2 functional regions: one binds with the target protein, and the other anchors to ALIX and ESCRT for delivery to multivesicular bodies (MVB) and lysosomes. In 2001, the first PROTAC was reported to recruit SCFβ-TRCP E3 and subsequently induce degradation of methionine aminopeptidase 2 (MetAp-2)52. In 2008, small molecule-based E3 recruitment ligands were invented, and great progress was made in PROTAC technology57. Studies indicated the feasibility of developing PROTACs that can enter cells relatively easily57. In 2013, mouse experiments provided the first demonstration that phospho-dependent PROTACs (PhosphoPROTACs) inhibit tumor growth *in vivo*58. Nonetheless, the peptidic E3 ligase ligands used in PROTACs have hindered their development into more mature chemical probes or therapeutic regimens, because peptidic E3 ligase ligands lead to high molecular weight of the entire PROTAC molecule, thus resulting in poor cell permeability59. In 2015, the novel PROTAC HaloPROTAC with incorporated small molecule VHL ligands was reported to successfully degrade HaloTag7 fusion proteins60. HaloTag7 is a modified bacterial dehalogenase that covalently reacts with hexyl chloride tags; HaloTag fusion proteins have been widely used to bioorthogonally label proteins *in vivo*. HaloPROTACs resulted in a 90% maximum degradation of GFP-HT7 with a low nanomolar half-maximum degradation concentration60. HaloPROTACs inspired the development of future PROTACs with more drug-like properties and have become useful chemical genetic tools61 as small molecule proteasome modulators. PROTAC is chimeric with these small molecules, and it forms a bifunctional small molecule compound that can link target proteins and E3 ubiquitin ligase in a ternary complex, thus resulting in target protein degradation through the ubiquitin-protease system21,22. Furthermore the development of small molecule-based PROTAC compounds with more drug-like properties has allowed for potent permeable PROTACs to be generated62,63. Studies have shown that a PROTAC against P300/CBP-associated factor (PCAF) and general control nonderepressible 5 (GCN5) effectively regulates the expression of multiple inflammatory mediators in macrophages and dendritic cells64. To date, a variety of PROTAC variants have been developed, thus laying a foundation for drug advancement. Homo-PROTACs have been developed for auto-targeting of both von Hippel-Lindau (VHL) and cereblon (CRBN)65,66. Additional insights have been gained in the structural basis and target selectivity of PROTACs. PROTACs can be designed to target various proteins of interest, because they are programmable67,68. Moreover, the PROTAC technology can escape the resistance mechanisms of inhibitors, including overexpression of target proteins and resistance mutations, by enabling modulation of both the enzymatic and non-enzymatic roles of proteins69. PROTAC technology has many advantages, such as low manufacturing cost, low drug dosage, excellent cell permeability, broad tissue distribution, and a strong ability to regulate intracellular targets70. PROTACs have already been applied for the degradation of various notable targets, including abelson murine leukemia viral oncogene homolog 1 (ABL)-breakpoint cluster region (BCR) in chronic myeloid leukemia, bromodomain proteins in multiple cancers, and androgen receptor in prostate cancer (details in **Table 1**)53,71,72. PD1/PD-L1 have not been selected as target proteins for study, although multiple protein targets have successfully been modulated with PROTAC technology. A study has reported that J22352, a highly selective HDAC6 inhibitor with PROTAC-like properties, decreases the immunosuppressive activity of PD-L1, thus restoring anti-tumor activity in glioblastoma117. View this table: [Table 1](http://www.cancerbiomed.org/content/17/3/583/T1) Table 1 Molecules that target disease-related proteins for proteasomal degradation ### Targeted lysosomal degradation of PD-L1 by PD-LYSO Although antibodies, small molecule inhibitors, and PROTACs have effects on immunotherapeutic target inhibition, the relatively low response rate and checkpoint blockade resistance have necessitated exploration of the molecular regulatory mechanisms of PD-L1. Studies have revealed the mechanisms that control PD-L1 transcriptional activation and post-translational modifications118–120. β transducin repeat containing protein (β-TrCP), cullin 3, and COP9 signaling body 5 (CSN5) control the degradation of PD-L1 through regulating PD-L1 ubiquitination121,122. Moreover, targeted blockade of PD-L1 transport from the endoplasmic reticulum to the Golgi apparatus triggers endoplasmic reticulum-related degradation of PD-L1123. Studies have found that chemokine-like factor (CKLF)-like MARVEL transmembrane domain-containing proteins 6 and 4 (CMTM6 and CMTM4) increase the stability of PD-L1 through downregulating ubiquitination-dependent degradation and lysosome-dependent proteolysis, thus enhancing the ability of tumor cells to suppress immune responses, and providing a new target for combinatorial immunotherapy37,124,125. In fact, the transport between recycling endosomes and lysosomes controls the fate of the PD-L1 protein37,126, although the exact mechanism of lysosomal-dependent degradation of PD-L1 is incompletely understood. Antitumor immunity is enhanced by inhibiting PD-L1, on the basis of the molecular regulation of PD-L1 in tumor cells. Our previous studies have shown that depletion of huntingtin-interacting protein 1-related protein (HIP1R) in tumor cells leads to significant upregulation of PD-L1, thus resulting in the suppression of T cell cytotoxicity24. Further research has shown that HIP1R is a regulator of PD-L1 lysosomal degradation that controls PD-L1 homeostasis. HIP1R physically interacts with PD-L1 and transports PD-L1 to lysosomes through a lysosomal targeting signal. HIP1R is an endocytic adaptor protein that contains homology domains responsible for the binding of clathrin, inositol lipids and F-actin127. HIP1R binds PD-L1 through its conserved C-terminal domain and uses an intrinsic sorting signal to deliver PD-L1 to lysosomes for degradation128. HIP1R targets PD-L1 for lysosomal degradation, thereby enhancing T cell-mediated cytotoxicity, and it is a natural regulator of lysosomal degradation. On the basis of the ‘binding–sorting’ model derived from the molecular roles of HIP1R, we have rationally designed the peptide PD-LYSO, incorporating the lysosome-sorting signal and the PD-L1-binding sequence of HIP1R, and used it to successfully deplete PD-L1 expression in tumor cells129. Other researchers have identified SA-49 as a novel regulator of PD-L1 expression from a series of novel aloperine derivatives. They have found that SA-49-induced microphthalmia transcription factor (MITF) translocation functions through activation of PKCα and subsequent suppression of GSK3β activity, thus increasing lysosome biogenesis and promoting translocation of PD-L1 to lysosomes for proteolysis126. In breast cancer, another study has identified a disintegrin and metalloproteinase 10 (ADAM10) and ADAM17 as enzymes mediating PD-L1 cleavage. The cleavage generates a free N-terminal fragment and a C-terminal fragment that remains associated with cells but is efficiently eliminated by lysosomal degradation130. Researchers are increasingly focusing on exploring more techniques for targeting proteins for lysosomal degradation, including endosome targeting chimeras (ENDTACs), lysosome targeting chimeras (LYTACs)24 (**Table 2**). View this table: [Table 2](http://www.cancerbiomed.org/content/17/3/583/T2) Table 2 Lysosomal targeting molecules reported in previous studies Thus, the discovery of HIP1R-mediated lysosomal degradation of PD-L1 has provided a potential new route for inhibiting PD-L1. PD-LYSO should be beneficial in the development and optimization of lysosomal targeting strategies as a crucial target for combinatorial immunotherapy. ## Palmitoylation blockade triggers degradation of PD-L1 and PD-1 The intracellular storage and redistribution of PD-L1 to cell membranes minimize the therapeutic benefits137. The cytoplasmic domain of PD-L1 is palmitoylated, and this lipid modification stabilizes PD-L1 by preventing its ubiquitination, thereby inhibiting lysosomal degradation24,138. Palmitoylation of proteins through linkage to 16-carbon fatty acid palmitate regulates protein localization and function139,140. Palmitate is usually associated with cysteine residues through thioester bonds, in a process that may be catalyzed by aspartic acid-histidine-histidine-cysteine (DHHC) palmitosyltransferase141. Palmitoylation is a reversible lipid modification of proteins that controls a variety of protein functions, such as transport, activity, stability, and membrane association142. Palmitoylation has been shown to regulate the transportation and function of multiple cancer-related proteins143,144. Research has indicated that palmitoylation plays an important role in the regulation of PD-L1 protein stability and trafficking and has identified the palmitoyltransferase ZDHHC3 (DHHC3) as the main acetyltransferase required for the palmitoylation of PD-L1145. Palmitoylation decreases the lysosomal degradation of PD-L1. The compound 2-bromopalmitate, a small-molecule inhibitor of palmitoylation, blocks the palmitoylation of PD-L1 and effectively induces lysosomal degradation of PD-L1 in tumor cells, thus enhancing the cytotoxicity of tumor-specific T cells24,138. The lack of specificity is a major challenge in targeting palmitoylation with existing palmitoylation inhibitors. Apart from the PD-L1-related adverse effects, 2-bromopalmitate might cause adverse effects related to its inhibitory effects on other palmitoylated proteins. PD-PALM, a PD-L1 palmitoylation inhibitor, has been designed to competitively inhibit PD-L1 palmitoylation. The application of PD-PALM decreases PD-L1 expression in tumor cells and enhances T cell activity24. Hence, inhibiting palmitoylation of PD-L1 may decrease PD-L1 expression on the cell membrane and deplete its storage capacity in recycling endosomes. Palmitoylation-based targeting methods may provide more powerful and long-lasting inhibitory effects because they inhibit PD-L1 protein levels throughout the cell and therefore may represent a promising therapeutic avenue toward enhancing tumor-specific immunity. ## Challenges and possibilities in future studies Because most targeted degradation mechanisms rely on intracellular binding and sorting, the cell-penetration ability of therapeutic molecules poses a major challenge. The peptidic nature of PD-LYSO and PD-PALM also makes *in vivo* stability an outstanding challenge, because peptides are generally prone to rapid degradation in the serum, through the action of various enzymes. To achieve a longer half-life, commonly used approaches, such as PEGylation, bovine serum albumin fusion, and Fc fusion, may significantly increase the molecular size and prohibit intracellular delivery. Moreover, potential immunogenic effects should be considered, because peptides may stimulate the generation of neutralizing antibodies *in vivo*. Because of these challenges, developing small molecules to mimic the conformation and functions of peptides is highly preferable and would represent a major step toward successful drug development. ## Acknowledgements This work was supported by the National Natural Science Foundation of China (Grant Nos. 81874050, 81572326, 81322036, 81421001, and 81902906), National Key R & D Program of China (Grant No. 2016YFC0906002), Startup Research Funding from Fudan University (Grant No. 2019XJ), and Jiangsu Province’s Medical Scientific Research Project (Grant No. H2019102). ## Footnotes * **Conflict of interest statement** No potential conflicts of interest are disclosed. * Received February 16, 2018. * Accepted April 30, 2020. * Copyright: © 2020, Cancer Biology & Medicine [https://creativecommons.org/licenses/by/4.0/](https://creativecommons.org/licenses/by/4.0/) This is an open access article distributed under the terms of the [Creative Commons Attribution License (CC BY) 4.0](https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution and reproduction in any medium, provided the original author and source are credited. ## References 1. 1. Topalian SL, Taube JM, Pardoll DM. Neoadjuvant checkpoint blockade for cancer immunotherapy. Science. 2020; 367. Doi: 10.1126/science.aax0182. In press. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1126/science.aax0182&link_type=DOI) 2. 2. Heyman B, Yang YP. New developments in immunotherapy for lymphoma. Cancer Biol Med. 2018; 15: 189–209. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiY2JtIjtzOjU6InJlc2lkIjtzOjg6IjE1LzMvMTg5IjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTcvMy81ODMuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 3. 3. Muir C, Menzies AM, Clifton-Bligh RJ, Tsang VHM. Thyroid toxicity following immune checkpoint inhibitor treatment in advanced cancer. Thyroid. 2020. Doi: 10.1089/thy.2020.0032. In press. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1089/thy.2020.0032&link_type=DOI) 4. 4. Park JG, Lee CR, Kim MG, Kim G, Shin HM, Jeon YH, et al. Kidney residency of VISTA-positive macrophages accelerates repair from ischemic injury. Kidney Int. 2019; 97: 980–94. 5. 5. Okoye IS, Xu L, Walker J, Elahi S. The glucocorticoids prednisone and dexamethasone differentially modulate T cell function in response to anti-PD-1 and anti-CTLA-4 immune checkpoint blockade. Cancer Immunol Immunother. 2020. Doi: 10.1007/s00262-020-02555-2. In press. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1007/s00262-020-02555-2&link_type=DOI) 6. 6. Zerdes I, Matikas A, Bergh J, Rassidakis GZ, Foukakis T. Genetic, transcriptional and post-translational regulation of the programmed death protein ligand 1 in cancer: biology and clinical correlations. Oncogene. 2018; 37: 4639–61. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 7. 7. Wang YT, Wang HB, Yao H, Li CS, Fang JY, Xu J. Regulation of PD-L1: emerging routes for targeting tumor immune evasion. Front Pharmacol. 2018; 9: 536. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.3389/fphar.2018.00536&link_type=DOI) 8. 8. Yao H, Wang HB, Li CS, Fang JY, Xu J. Cancer cell-intrinsic PD-1 and implications in combinatorial immunotherapy. Front Immunol. 2018; 9: 1774. 9. 9. Shimizu K, Sugiura D, Okazaki IM, Maruhashi T, Takegami Y, Cheng CY, et al. PD-1 imposes qualitative control of cellular transcriptomes in response to T cell activation. Mol Cell. 2020; 77: 937–50. 10. 10. Wei YH, Du Q, Jiang XY, Li L, Li T, Li MQ, et al. Efficacy and safety of combination immunotherapy for malignant solid tumors: a systematic review and meta-analysis. Crit Rev Oncol Hematol. 2019; 138: 178–89. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 11. 11. Dougan M, Pietropaolo M. Time to dissect the autoimmune etiology of cancer antibody immunotherapy. J Clin Invest. 2020; 130: 51–61. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 12. 12. Moser JC, Hu-Lieskovan S. Mechanisms of resistance to PD-1 checkpoint blockade. Drugs. 2020; 80: 459–65. 13. 13. Friedman CF, Snyder A. Atypical autoimmune adverse effects with checkpoint blockade therapies. Ann Oncol. 2017; 28: 206–7. 14. 14. Wang HF, Fu C, Du J, Wang HS, He R, Yin XF, et al. Enhanced histone H3 acetylation of the PD-L1 promoter *via* the COP1/c-Jun/HDAC3 axis is required for PD-L1 expression in drug-resistant cancer cells. J Exp Clin Cancer Res. 2020; 39: 29. 15. 15. Zhang R, Zhu ZY, Lv HY, Li FT, Sun SQ, Li J, et al. Immune checkpoint blockade mediated by a small-molecule nanoinhibitor targeting the PD-1/PD-L1 pathway synergizes with photodynamic therapy to elicit antitumor immunity and antimetastatic effects on breast cancer. Small. 2019; 15: e1903881. 16. 16. Hu ZP, Yu PF, Du GY, Wang WY, Zhu HB, Li N, et al. PCC0208025 (BMS202), a small molecule inhibitor of PD-L1, produces an antitumor effect in B16-F10 melanoma-bearing mice. PLoS One. 2020; 15: e0228339. 17. 17. Adjei AA. What is the right dose? The elusive optimal biologic dose in phase I clinical trials. J Clin Oncol. 2006; 24: 4054–5. 18. 18. Soutschek J, Akinc A, Bramlage B, Charisse K, Constien R, Donoghue M, et al. Therapeutic silencing of an endogenous gene by systemic administration of modified siRNAs. Nature. 2004; 432: 173–8. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nature03121&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=15538359&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000225020200035&link_type=ISI) 19. 19. Bumcrot D, Manoharan M, Koteliansky V, Sah DW. RNAi therapeutics: a potential new class of pharmaceutical drugs. Nat Chem Biol. 2006; 2: 711–9. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nchembio839&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=17108989&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000242168800015&link_type=ISI) 20. 20. Tokatlian T, Segura T. siRNA applications in nanomedicine. Wiley Interdiscip Rev Nanomed Nanobiotechnol. 2010; 2: 305–15. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/wnan.81&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=20135697&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 21. 21. Bashraheel SS, Domling A, Goda SK. Update on targeted cancer therapies, single or in combination, and their fine tuning for precision medicine. Biomed Pharmacother. 2020; 125: 110009. 22. 22. Zhang XH, Lee HC, Shirazi F, Baladandayuthapani V, Lin H, Kuiatse I, et al. Protein targeting chimeric molecules specific for bromodomain and extra-terminal motif family proteins are active against pre-clinical models of multiple myeloma. Leukemia. 2018; 32: 2224–39. 23. 23. Li JY, Yang Y, Yu Y, Li QB, Tan GX, Wang YY, et al. LAPONITE® nanoplatform functionalized with histidine modified oligomeric hyaluronic acid as an effective vehicle for the anticancer drug methotrexate. J Mater Chem B. 2018; 6: 5011–20. 24. 24. Yao H, Lan J, Li CS, Shi HB, Brosseau JP, Wang HB, et al. Author correction: inhibiting PD-L1 palmitoylation enhances T-cell immune responses against tumours. Nat Biomed Eng. 2019; 3: 414. 25. 25. Vaddepally RK, Kharel P, Pandey R, Garje R, Chandra AB. Review of indications of FDA-approved immune checkpoint inhibitors per NCCN guidelines with the level of evidence. Cancers (Basel). 2020; 12: 738. 26. 26. Ali MHM, Toor SM, Rakib F, Mall R, Ullah E, Mroue K, et al. Investigation of the effect of PD-L1 blockade on triple negative breast cancer cells using fourier transform infrared spectroscopy. Vaccines (Basel). 2019; 7: 109. 27. 27. Dong H, Strome SE, Salomao DR, Tamura H, Hirano F, Flies DB, et al. Tumor-associated B7-H1 promotes T-cell apoptosis: a potential mechanism of immune evasion. Nat Med. 2002; 8: 793–800. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nm730&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=12091876&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000177200900024&link_type=ISI) 28. 28. Jefferis R, Lund J. Interaction sites on human IgG-Fc for FcgammaR: current models. Immunol Lett. 2002; 82: 57–65. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/S0165-2478(02)00019-6&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=12008035&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000176059200009&link_type=ISI) 29. 29. Brahmer JR, Drake CG, Wollner I, Powderly JD, Picus J, Sharfman WH, et al. Phase I study of single-agent anti-programmed death-1 (MDX-1106) in refractory solid tumors: safety, clinical activity, pharmacodynamics, and immunologic correlates. J Clin Oncol. 2010; 28: 3167–75. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamNvIjtzOjU6InJlc2lkIjtzOjEwOiIyOC8xOS8zMTY3IjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTcvMy81ODMuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 30. 30. Hamid O, Robert C, Daud A, Hodi FS, Hwu WJ, Kefford R, et al. Safety and tumor responses with lambrolizumab (anti-PD-1) in melanoma. N Engl J Med. 2013; 369: 134–44. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1056/NEJMoa1305133&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23724846&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000321567300009&link_type=ISI) 31. 31. Topalian SL, Hodi FS, Brahmer JR, Gettinger SN, Smith DC, McDermott DF, et al. Safety, activity, and immune correlates of anti-PD-1 antibody in cancer. N Engl J Med. 2012; 366: 2443–54. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1056/NEJMoa1200690&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=22658127&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000305747000004&link_type=ISI) 32. 32. Tan SG, Zhang H, Chai Y, Song H, Tong Z, Wang QH, et al. An unexpected N-terminal loop in PD-1 dominates binding by nivolumab. Nat Commun. 2017; 8: 14369. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/ncomms14369&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=28165004&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 33. 33. Rizvi NA, Hellmann MD, Brahmer JR, Juergens RA, Borghaei H, Gettinger S, et al. Nivolumab in combination with platinum-based doublet chemotherapy for first-line treatment of advanced non-small-cell lung cancer. J Clin Oncol. 2016; 34: 2969–79. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamNvIjtzOjU6InJlc2lkIjtzOjEwOiIzNC8yNS8yOTY5IjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTcvMy81ODMuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 34. 34. Borghaei H, Paz-Ares L, Horn L, Spigel DR, Steins M, Ready NE, et al. Nivolumab versus docetaxel in advanced nonsquamous non-small-cell lung cancer. N Engl J Med. 2015; 373: 1627–39. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1056/NEJMoa1507643&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=26412456&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 35. 35. Fiala O, Sorejs O, Sustr J, Kucera R, Topolcan O, Finek J. Immune-related adverse effects and outcome of patients with cancer treated with immune checkpoint inhibitors. Anticancer Res. 2020; 40: 1219–27. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MTA6ImFudGljYW5yZXMiO3M6NToicmVzaWQiO3M6OToiNDAvMy8xMjE5IjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTcvMy81ODMuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 36. 36. Baroudjian B, Arangalage D, Cuzzubbo S, Hervier B, Lebbé C, Lorillon G, et al. Management of immune-related adverse events resulting from immune checkpoint blockade. Expert Rev Anticancer Ther. 2019; 19: 209–22. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 37. 37. Burr ML, Sparbier CE, Chan YC, Williamson JC, Woods K, Beavis PA, et al. CMTM6 maintains the expression of PD-L1 and regulates anti-tumour immunity. Nature. 2017; 549: 101–5. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 38. 38. Xie FT, Xu MX, Lu J, Mao LX, Wang SJ. The role of exosomal PD-L1 in tumor progression and immunotherapy. Mol Cancer. 2019; 18: 146. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 39. 39. Seo N, Akiyoshi K, Shiku H. Exosome-mediated regulation of tumor immunology. Cancer Sci. 2018; 109: 2998–3004. 40. 40. Chen G, Huang AC, Zhang W, Zhang G, Wu M, Xu W, et al. Exosomal PD-L1 contributes to immunosuppression and is associated with anti-PD-1 response. Nature. 2018; 560: 382–6. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/s41586-018-0392-8&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=30089911&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 41. 41. Zhai WJ, Zhou XM, Du JF, Gao YF. In vitro assay for the development of small molecule inhibitors targeting PD-1/PD-L1. Meth Enzymol. 2019; 629: 361–81. 42. 42. Li JD, Van Valkenburgh J, Hong XF, Conti PS, Zhang XZ, Chen K. Small molecules as theranostic agents in cancer immunology. Theranostics. 2019; 9: 7849–71. 43. 43. Collier AM, Nemtsova Y, Kuber N, Banach-Petrosky W, Modak A, Sleat DE, et al. Lysosomal protein thermal stability does not correlate with cellular half-life: global observations and a case study of tripeptidyl-peptidase 1. Biochem J. 2020; 477: 727–45. 44. 44. Li CL, Zhang NP, Zhou DJ, Ding C, Jin YQ, Cui XY, et al. Peptide blocking of PD-1/PD-L1 interaction for cancer immunotherapy. Cancer Immunol Res. 2018; 6: 178–88. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NjoiY2FuaW1tIjtzOjU6InJlc2lkIjtzOjc6IjYvMi8xNzgiO3M6NDoiYXRvbSI7czoxODoiL2NibS8xNy8zLzU4My5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 45. 45. Sasikumar PG, Ramachandra RK, Adurthi S, Dhudashiya AA, Vadlamani S, Vemula K, et al. A rationally designed peptide antagonist of the PD-1 signaling pathway as an immunomodulatory agent for cancer therapy. Mol Cancer Ther. 2019; 18: 1081–91. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MTA6Im1vbGNhbnRoZXIiO3M6NToicmVzaWQiO3M6OToiMTgvNi8xMDgxIjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTcvMy81ODMuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 46. 46. Musielak B, Kocik J, Skalniak L, Magiera-Mularz K, Sala D, Czub M, et al. CA-170 - a potent small-molecule PD-L1 inhibitor or not? Molecules. 2019; 24: 2804. 47. 47.United States patent application, US20130022629A1 accessed on 20 November 2019. 48. 48.United States patent application, US9850225 accessed on 20 November 2019. 49. 49. Chang HN, Liu BY, Qi YK, Zhou Y, Chen YP, Pan KM, et al. Blocking of the PD-1/PD-L1 interaction by a D-peptide antagonist for cancer immunotherapy. Angew Chem Int Ed Engl. 2015; 54: 11760–4. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/anie.201506225&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 50. 50. Kondo M, Kikumoto H, Osimitz TG, Cohen SM, Lake BG, Yamada T. An evaluation of the human relevance of the liver tumors observed in female mice treated with permethrin based on mode of action. Toxicol Sci. 2020; 175: 50–63. 51. 51. do Amaral DF, Guerra V, Motta AGC, de Melo E Silva D, Rocha TL. Ecotoxicity of nanomaterials in amphibians: a critical review. Sci Total Environ. 2019; 686: 332–44. 52. 52. Sakamoto KM, Kim KB, Kumagai A, Mercurio F, Crews CM, Deshaies RJ. Protacs: chimeric molecules that target proteins to the Skp1-Cullin-F box complex for ubiquitination and degradation. Proc Natl Acad Sci USA. 2001; 98: 8554–9. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NDoicG5hcyI7czo1OiJyZXNpZCI7czoxMDoiOTgvMTUvODU1NCI7czo0OiJhdG9tIjtzOjE4OiIvY2JtLzE3LzMvNTgzLmF0b20iO31zOjg6ImZyYWdtZW50IjtzOjA6IiI7fQ==) 53. 53. Bondeson DP, Mares A, Smith IE, Ko E, Campos S, Miah AH, et al. Catalytic in vivo protein knockdown by small-molecule PROTACs. Nat Chem Biol. 2015; 11: 611–7. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nchembio.1858&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=26075522&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 54. 54. Fisher SL, Phillips AJ. Targeted protein degradation and the enzymology of degraders. Curr Opin Chem Biol. 2018; 44: 47–55. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=29885948&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 55. 55. Smalley JP, Adams GE, Millard CJ, Song Y, Norris JKS, Schwabe JWR, et al. PROTAC-mediated degradation of class I histone deacetylase enzymes in corepressor complexes. Chem Commun (Camb). 2020; 56: 4476–9. 56. 56. Wan YC, Yan CX, Gao H, Liu TT. Small-molecule PROTACs: novel agents for cancer therapy. Future Med Chem. 2020; 12: 915–38. 57. 57. Schneekloth AR, Pucheault M, Tae HS, Crews CM. Targeted intracellular protein degradation induced by a small molecule: en route to chemical proteomics. Bioorg Med Chem Lett. 2008; 18: 5904–8. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.bmcl.2008.07.114&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=18752944&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 58. 58. Hines J, Gough JD, Corson TW, Crews CM. Posttranslational protein knockdown coupled to receptor tyrosine kinase activation with phosphoPROTACs. Proc Natl Acad Sci USA. 2013; 110: 8942–7. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NDoicG5hcyI7czo1OiJyZXNpZCI7czoxMToiMTEwLzIyLzg5NDIiO3M6NDoiYXRvbSI7czoxODoiL2NibS8xNy8zLzU4My5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 59. 59. Jiang YH, Deng QW, Zhao H, Xie MS, Chen LJ, Yin F, et al. Development of stabilized peptide-based PROTACs against estrogen receptor . ACS Chem Biol. 2018; 13: 628–35. 60. 60. Buckley DL, Raina K, Darricarrere N, Hines J, Gustafson JL, Smith IE, et al. HaloPROTACS: use of small molecule PROTACs to induce degradation of halotag fusion proteins. ACS Chem Biol. 2015; 10: 1831–7. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1021/acschembio.5b00442&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=26070106&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 61. 61. Tovell H, Testa A, Maniaci C, Zhou H, Prescott AR, Macartney T, et al. Rapid and reversible knockdown of endogenously tagged endosomal proteins *via* an optimized HaloPROTAC degrader. ACS Chem Biol. 2019; 14: 882–92. 62. 62. Zorba A, Nguyen C, Xu Y, Starr J, Borzilleri K, Smith J, et al. Delineating the role of cooperativity in the design of potent PROTACs for BTK. Proc Natl Acad Sci USA. 2018; 115: E7285–E92. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NDoicG5hcyI7czo1OiJyZXNpZCI7czoxMjoiMTE1LzMxL0U3Mjg1IjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTcvMy81ODMuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 63. 63. Jiang F, Wei QY, Li HL, Li HM, Cui Y, Ma Y, et al. Discovery of novel small molecule induced selective degradation of the bromodomain and extra-terminal (BET) bromodomain protein BRD4 and BRD2 with cellular potencies. Bioorg Med Chem. 2020; 28: 115181. 64. 64. Bassi ZI, Fillmore MC, Miah AH, Chapman TD, Maller C, Roberts EJ, et al. Modulating PCAF/GCN5 immune cell function through a PROTAC approach. ACS Chem Biol. 2018; 13: 2862–7. 65. 65. Maniaci C, Hughes SJ, Testa A, Chen W, Lamont DJ, Rocha S, et al. Homo-PROTACs: bivalent small-molecule dimerizers of the VHL E3 ubiquitin ligase to induce self-degradation. Nat Commun. 2017; 8: 830. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/s41467-017-00954-1&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=29018234&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 66. 66. Steinebach C, Lindner S, Udeshi ND, Mani DC, Kehm H, Köpff S, et al. Homo-PROTACs for the chemical knockdown of cereblon. ACS Chem Biol. 2018; 13: 2771–82. 67. 67. Bondeson DP, Smith BE, Burslem GM, Buhimschi AD, Hines J, Jaime-Figueroa S, et al. Lessons in PROTAC design from selective degradation with a promiscuous warhead. Cell Chem Biol. 2018; 25: 78–87.e5. 68. 68. Gadd MS, Testa A, Lucas X, Chan KH, Chen W, Lamont DJ, et al. Structural basis of PROTAC cooperative recognition for selective protein degradation. Nat Chem Biol. 2017; 13: 514–21. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nchembio.2329&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=28288108&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 69. 69. Clague MJ, Heride C, Urbé S. The demographics of the ubiquitin system. Trends Cell Biol. 2015; 25: 417–26. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.tcb.2015.03.002&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=25906909&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 70. 70. Zhou B, Hu JT, Xu FM, Chen Z, Bai LC, Fernandez-Salas E, et al. Discovery of a small-molecule degrader of bromodomain and extra-terminal (BET) proteins with picomolar cellular potencies and capable of achieving tumor regression. J Med Chem. 2018; 61: 462–81. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1021/acs.jmedchem.6b01816&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=28339196&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 71. 71. Burslem GM, Smith BE, Lai AC, Jaime-Figueroa S, McQuaid DC, Bondeson DP, et al. The advantages of targeted protein degradation over inhibition: an RTK case study. Cell Chem Biol. 2018; 25: 67–77. 72. 72. Burslem GM, Schultz AR, Bondeson DP, Eide CA, Savage Stevens SL, Druker BJ, et al. Targeting BCR-ABL1 in chronic myeloid leukemia by PROTAC-mediated targeted protein degradation. Cancer Res. 2019; 79: 4744–53. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1158/1538-7445.SABCS18-4744&link_type=DOI) 73. 73. Zhao QJ, Ren CW, Liu LY, Chen JJ, Shao YB, Sun N, et al. Discovery of SIAIS178 as an effective BCR-ABL degrader by recruiting Von Hippel-Lindau (VHL) E3 ubiquitin ligase. J Med Chem. 2019; 62: 9281–98. 74. 74. Lai AC, Toure M, Hellerschmied D, Salami J, Jaime-Figueroa S, Ko E, et al. Modular PROTAC design for the degradation of oncogenic BCR-ABL. Angew Chem Int Ed Engl. 2016; 55: 807–10. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/anie.201507634&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=26593377&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 75. 75. Zhang H, Zhao HY, Xi XX, Liu YJ, Xin M, Mao S, et al. Discovery of potent epidermal growth factor receptor (EGFR) degraders by proteolysis targeting chimera (PROTAC). Eur J Med Chem. 2020; 189: 112061. 76. 76. Kargbo RB. Treatment of cancer and alzheimer’s disease by PROTAC degradation of EGFR. ACS Med Chem Lett. 2019; 10: 1098–9. 77. 77. Crew AP, Raina K, Dong H, Qian Y, Wang J, Vigil D, et al. Identification and characterization of Von Hippel-Lindau-recruiting proteolysis targeting chimeras (PROTACs) of TANK-binding kinase 1. J Med Chem. 2018; 61: 583–98. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1021/acs.jmedchem.7b00635&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=28692295&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 78. 78. Rana S, Bendjennat M, Kour S, King HM, Kizhake S, Zahid M, et al. Selective degradation of CDK6 by a palbociclib based PROTAC. Bioorg Med Chem Lett. 2019; 29: 1375–9. 79. 79. Bian JL, Ren J, Li YR, Wang JB, Xu X, Feng YF, et al. Discovery of Wogonin-based PROTACs against CDK9 and capable of achieving antitumor activity. Bioorg Chem. 2018; 81: 373–81. 80. 80. Robb CM, Contreras JI, Kour S, Taylor MA, Abid M, Sonawane YA, et al. Chemically induced degradation of CDK9 by a proteolysis targeting chimera (PROTAC). Chem. Commun (Camb). 2017; 53: 7577–80. 81. 81. Zhang C, Han XR, Yang X, Jiang B, Liu J, Xiong Y, et al. Proteolysis targeting chimeras (PROTACs) of anaplastic lymphoma kinase (ALK). Eur J Med Chem. 2018; 151: 304–14. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.ejmech.2018.03.071&link_type=DOI) 82. 82. Kang CH, Lee DH, Lee CO, Du Ha J, Park CH, Hwang JY. Induced protein degradation of anaplastic lymphoma kinase (ALK) by proteolysis targeting chimera (PROTAC). Biochem Biophys Res Commun. 2018; 505: 542–7. 83. 83. Tovell H, Testa A, Zhou H, Shpiro N, Crafter C, Ciulli A, et al. Design and characterization of SGK3-PROTAC1, an isoform specific SGK3 kinase PROTAC degrader. ACS Chem Biol. 2019; 14: 2024–34. 84. 84. Chen H, Chen FH, Liu NN, Wang XY, Gou SH. Chemically induced degradation of CK2 by proteolysis targeting chimeras based on a ubiquitin-proteasome pathway. Bioorg Chem. 2018; 81: 536–44. 85. 85. Vollmer S, Cunoosamy D, Lv HF, Feng HX, Li X, Nan ZY, et al. Design, synthesis, and biological evaluation of MEK PROTACs. J Med Chem. 2020; 63: 157–62. 86. 86. Lebraud H, Wright DJ, Johnson CN, Heightman TD. Protein degradation by in-cell self-assembly of proteolysis targeting chimeras. ACS Cent Sci. 2016; 2: 927–34. 87. 87. Burslem GM, Song J, Chen X, Hines J, Crews CM. Enhancing antiproliferative activity and selectivity of a FLT-3 inhibitor by proteolysis targeting chimera conversion. J Am Chem Soc. 2018; 140: 16428–32. 88. 88. Li WL, Gao CM, Zhao L, Yuan ZG, Chen YZ, Jiang YY. Phthalimide conjugations for the degradation of oncogenic PI3K. Eur J Med Chem. 2018; 151: 237–47. 89. 89. You I, Erickson EC, Donovan KA, Eleuteri NA, Fischer ES, Gray NS, et al. Discovery of an AKT degrader with prolonged inhibition of downstream signaling. Cell Chem Biol. 2020; 27: 66–73.e7. 90. 90. Buhimschi AD, Armstrong HA, Toure M, Jaime-Figueroa S, Chen TL, Lehman AM, et al. Targeting the C481S ibrutinib-resistance mutation in Bruton’s tyrosine kinase using PROTAC-mediated degradation. Biochemistry. 2018; 57: 3564–75. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1021/acs.biochem.8b00391&link_type=DOI) 91. 91. Sun YH, Ding N, Song YQ, Yang ZM, Liu WL, Zhu J, et al. Degradation of Bruton’s tyrosine kinase mutants by PROTACs for potential treatment of ibrutinib-resistant non-Hodgkin lymphomas. Leukemia. 2019; 33: 2105–10. 92. 92. Sun YH, Zhao XW, Ding N, Gao HY, Wu Y, Yang YQ, et al. PROTAC-induced BTK degradation as a novel therapy for mutated BTK C481S induced ibrutinib-resistant B-cell malignancies. Cell Res. 2018; 28: 779–81. 93. 93. Cromm PM, Samarasinghe KTG, Hines J, Crews CM. Addressing kinase-independent functions of fak *via* PROTAC-mediated degradation. J Am Chem Soc. 2018; 140: 17019–26. 94. 94. Nunes J, McGonagle GA, Eden J, Kiritharan G, Touzet M, Lewell X, et al. Targeting IRAK4 for degradation with PROTACs. ACS Med Chem Lett. 2019; 10: 1081–5. 95. 95. Qin C, Hu Y, Zhou B, Fernandez-Salas E, Yang CY, Liu L, et al. Discovery of QCA570 as an exceptionally potent and efficacious proteolysis targeting chimera (PROTAC) degrader of the bromodomain and extra-terminal (BET) proteins capable of inducing complete and durable tumor regression. J Med Chem. 2018; 61: 6685–704. 96. 96. Zengerle M, Chan KH, Ciulli A. Selective small molecule induced degradation of the BET bromodomain protein BRD4. ACS Chem Biol. 2015; 10: 1770–7. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1021/acschembio.5b00216&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=26035625&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 97. 97. Lu J, Qian YM, Altieri M, Dong HQ, Wang J, Raina K, et al. Hijacking the E3 ubiquitin ligase cereblon to efficiently target BRD4. Chem Biol. 2015; 22: 755–63. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.chembiol.2015.05.009&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=26051217&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 98. 98. Hines J, Lartigue S, Dong H, Qian Y, Crews CM. MDM2-recruiting PROTAC offers superior, synergistic antiproliferative activity *via* simultaneous degradation of BRD4 and stabilization of p53. Cancer Res. 2019; 79: 251–62. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NjoiY2FucmVzIjtzOjU6InJlc2lkIjtzOjg6Ijc5LzEvMjUxIjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTcvMy81ODMuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 99. 99. Zoppi V, Hughes SJ, Maniaci C, Testa A, Gmaschitz T, Wieshofer C, et al. Iterative design and optimization of initially inactive proteolysis targeting chimeras (PROTACs) identify VZ185 as a potent, fast, and selective von Hippel-Lindau (VHL) based dual degrader probe of BRD9 and BRD7. J Med Chem. 2019; 62: 699–726. 100.100. Han X, Wang C, Qin C, Xiang WG, Fernandez-Salas E, Yang CY, et al. Discovery of ARD-69 as a highly potent proteolysis targeting chimera (PROTAC) degrader of androgen receptor (AR) for the treatment of prostate cancer. J Med Chem. 2019; 62: 941–64. 101.101. Wang Y, Jiang XY, Feng F, Liu WY, Sun H. Degradation of proteins by PROTACs and other strategies. Acta Pharm Sin B. 2020; 10: 207–38. 102.102. Salami J, Alabi S, Willard RR, Vitale NJ, Wang J, Dong H, et al. Androgen receptor degradation by the proteolysis-targeting chimera ARCC-4 outperforms enzalutamide in cellular models of prostate cancer drug resistance. Commun Biol. 2018; 1: 100. Doi: 10.1038/s42003-018-0105-8. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/s42003-018-0105-8&link_type=DOI) 103.103. Hu JT, Hu B, Wang ML, Xu FM, Miao B, Yang CY, et al. Discovery of ERD-308 as a highly potent proteolysis targeting chimera (PROTAC) degrader of estrogen receptor (ER). J Med Chem. 2019; 62: 1420–42. 104.104. Wang L, Guillen VS, Sharma N, Flessa K, Min J, Carlson KE, et al. New class of selective estrogen receptor degraders (SERDs): expanding the toolbox of PROTAC degrons. ACS Med Chem Lett. 2018; 9: 803–8. 105.105. Gechijian LN, Buckley DL, Lawlor MA, Reyes JM, Paulk J, Ott CJ, et al. Functional TRIM24 degrader *via* conjugation of ineffectual bromodomain and VHL ligands. Nat Chem Biol. 2018; 14: 405–12. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/s41589-018-0010-y&link_type=DOI) 106.106. Sakamoto KM, Kim KB, Verma R, Ransick A, Stein B, Crews CM, et al. Development of protacs to target cancer-promoting proteins for ubiquitination and degradation. Mol Cell Proteomics. 2003; 2: 1350–8. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NjoibWNwcm90IjtzOjU6InJlc2lkIjtzOjk6IjIvMTIvMTM1MCI7czo0OiJhdG9tIjtzOjE4OiIvY2JtLzE3LzMvNTgzLmF0b20iO31zOjg6ImZyYWdtZW50IjtzOjA6IiI7fQ==) 107.107. Zhang X, Thummuri D, He YH, Liu XG, Zhang PY, Zhou DH, et al. Utilizing PROTAC technology to address the on-target platelet toxicity associated with inhibition of BCL-X. Chem Commun (Camb). 2019; 55: 14765–8. 108.108. Schiedel M, Herp D, Hammelmann S, Swyter S, Lehotzky A, Robaa D, et al. Chemically induced degradation of sirtuin 2 (Sirt2) by a proteolysis targeting chimera (PROTAC) based on sirtuin rearranging ligands (SirReals). J Med Chem. 2018; 61: 482–91. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1021/acs.jmedchem.6b01872&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=28379698&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 109.109. An ZX, Lv WX, Su S, Wu W, Rao Y. Developing potent PROTACs tools for selective degradation of HDAC6 protein. Protein Cell. 2019; 10: 606–9. 110.110. Hsu JH, Rasmusson T, Robinson J, Pachl F, Read J, Kawatkar S, et al. EED-targeted PROTACs degrade EED, EZH2, and SUZ12 in the PRC2 complex. Cell Chem Biol. 2020; 27: 41–6. 111.111. Lu MC, Liu T, Jiao Q, Ji JN, Tao MM, Liu YJ, et al. Discovery of a Keap1-dependent peptide PROTAC to knockdown Tau by ubiquitination-proteasome degradation pathway. Eur J Med Chem. 2018; 146: 251–9. 112.112. Farnaby W, Koegl M, Roy MJ, Whitworth C, Diers E, Trainor N, et al. BAF complex vulnerabilities in cancer demonstrated *via* structure-based PROTAC design. Nat Chem Biol. 2019; 15: 672–80. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/s41589-019-0294-6&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=31178587&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 113.113. Zhao QY, Lan TL, Su S, Rao Y. Induction of apoptosis in MDA-MB-231 breast cancer cells by a PARP1-targeting PROTAC small molecule. Chem Commun (Camb) 2019; 55: 369–72. 114.114. Zhou HB, Bai LC, Xu RQ, Zhao YJ, Chen JY, McEachern D, et al. Structure-based discovery of SD-36 as a potent, selective, and efficacious PROTAC degrader of STAT3 protein. J Med Chem. 2019; 62: 11280–300. 115.115. Papatzimas JW, Gorobets E, Maity R, Muniyat MI, MacCallum JL, Neri P, et al. From inhibition to degradation: targeting the antiapoptotic protein myeloid cell leukemia 1 (MCL1). J Med Chem. 2019; 62: 5522–40. 116.116. Li YB, Yang JL, Aguilar A, McEachern D, Przybranowski S, Liu L, et al. Discovery of MD-224 as a first-in-class, highly potent, and efficacious proteolysis targeting chimera murine double minute 2 degrader capable of achieving complete and durable tumor regression. J Med Chem. 2019; 62: 448–66. 117.117. Liu JR, Yu CW, Hung PY, Hsin LW, Chern JW. High-selective HDAC6 inhibitor promotes HDAC6 degradation following autophagy modulation and enhanced antitumor immunity in glioblastoma. Biochem Pharmacol. 2019; 163: 458–71. 118.118. Dorand RD, Nthale J, Myers JT, Barkauskas DS, Avril S, Chirieleison SM, et al. Cdk5 disruption attenuates tumor PD-L1 expression and promotes antitumor immunity. Science. 2016; 353: 399–403. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6Mzoic2NpIjtzOjU6InJlc2lkIjtzOjEyOiIzNTMvNjI5Ny8zOTkiO3M6NDoiYXRvbSI7czoxODoiL2NibS8xNy8zLzU4My5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 119.119. Kataoka K, Shiraishi Y, Takeda Y, Sakata S, Matsumoto M, Nagano S, et al. Aberrant PD-L1 expression through 3-UTR disruption in multiple cancers. Nature. 2016; 534: 402–6. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nature18294&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=27281199&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 120.120. Kortlever RM, Sodir NM, Wilson CH, Burkhart DL, Pellegrinet L, Brown Swigart L, et al. Myc cooperates with ras by programming inflammation and immune suppression. Cell. 2017; 171: 1301–15.e14. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.cell.2017.11.013&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=29195074&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 121.121. Zhang JF, Bu X, Wang HZ, Zhu YS, Geng Y, Nihira NT, et al. Author correction: cyclin D-CDK4 kinase destabilizes PD-L1 *via* cullin 3-SPOP to control cancer immune surveillance. Nature. 2019; 571: E10. 122.122. Lim SO, Li CW, Xia W, Cha JH, Chan LC, Wu Y, et al. Deubiquitination and stabilization of PD-L1 by CSN5. Cancer Cell. 2016; 30: 925–39. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.ccell.2016.10.010&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=27866850&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 123.123. Cha JH, Yang WH, Xia W, Wei Y, Chan LC, Lim SO, et al. Metformin promotes antitumor immunity *via* endoplasmic-reticulum-associated degradation of PD-L1. Mol Cell. 2018; 71: 606–20.e7. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.molcel.2018.07.030&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 124.124. Mezzadra R, Sun C, Jae LT, Gomez-Eerland R, de Vries E, Wu W, et al. Identification of CMTM6 and CMTM4 as PD-L1 protein regulators. Nature. 2017; 549: 106–10. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nature23669&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=28813410&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 125.125. Koh YW, Han JH, Haam S, Jung J, Lee HW. Increased CMTM6 can predict the clinical response to PD-1 inhibitors in non-small cell lung cancer patients. Oncoimmunology. 2019; 8: e1629261. 126.126. Zhang N, Dou YY, Liu L, Zhang X, Liu XJ, Zeng QX, et al. SA-49, a novel aloperine derivative, induces MITF-dependent lysosomal degradation of PD-L1. EBioMedicine. 2019; 40: 151–62. 127.127. Peng L, Yang Q, Xu XX, Du YL, Wu Y, Shi XF, et al. Huntingtin-interacting protein 1-related protein plays a critical role in dendritic development and excitatory synapse formation in hippocampal neurons. Front Mol Neurosci. 2017; 10: 186. 128.128. Skruzny M, Brach T, Ciuffa R, Rybina S, Wachsmuth M, Kaksonen M. Molecular basis for coupling the plasma membrane to the actin cytoskeleton during clathrin-mediated endocytosis. Proc Natl Acad Sci USA. 2012; 109: E2533–42. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NDoicG5hcyI7czo1OiJyZXNpZCI7czoxMjoiMTA5LzM4L0UyNTMzIjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTcvMy81ODMuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 129.129. Wang HB, Yao H, Li CS, Shi HB, Lan J, Li ZL, et al. HIP1R targets PD-L1 to lysosomal degradation to alter T cell-mediated cytotoxicity. Nat Chem Biol. 2019; 15: 42–50. 130.130. Romero Y, Wise R, Zolkiewska A. Proteolytic processing of PD-L1 by ADAM proteases in breast cancer cells. Cancer Immunol Immunother. 2020; 69: 43–55. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 131.131. Dvela-Levitt M, Kost-Alimova M, Emani M, Kohnert E, Thompson R, Sidhom EH, et al. Small molecule targets TMED9 and promotes lysosomal degradation to reverse proteinopathy. Cell. 2019; 178: 521–35.e23. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.cell.2019.07.002&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=31348885&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 132.132. Banik SM, Pedram K, Wisnovsky S, Riley NM, Bertozzi CR. Lysosometargeting chimeras (LYTACs) for the degradation of secreted and membrane proteins. ChemRxiv. 2019; 1–68. 133.133. Hong JH, Kaustov L, Coyaud E, Srikumar T, Wan J, Arrowsmith C, et al. KCMF1 (potassium channel modulatory factor 1) links RAD6 to UBR4 (ubiquitin N-recognin domain-containing E3 ligase 4) and lysosome-mediated degradation. Mol Cell Proteomics. 2015; 14: 674–85. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NjoibWNwcm90IjtzOjU6InJlc2lkIjtzOjg6IjE0LzMvNjc0IjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTcvMy81ODMuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 134.134. Dominguez-Brauer C, Hao Z, Elia AJ, Fortin JM, Nechanitzky R, Brauer PM, et al. Mule regulates the intestinal stem cell niche *via* the Wnt pathway and targets EphB3 for proteasomal and lysosomal degradation. Cell Stem Cell. 2016; 19: 205–16. 135.135. Zhou YF, Wang J, Deng MF, Chi B, Wei N, Chen JG, et al. The peptide-directed lysosomal degradation of CDK5 exerts therapeutic effects against stroke. Aging Dis. 2019; 10: 1140–5. 136.136. Fan X, Jin WY, Lu J, Wang J, Wang YT. Rapid and reversible knockdown of endogenous proteins by peptide-directed lysosomal degradation. Nat Neurosci. 2014; 17: 471–80. 137.137. Monypenny J, Milewicz H, Flores-Borja F, Weitsman G, Cheung A, Chowdhury R, et al. ALIX regulates tumor-mediated immunosuppression by controlling EGFR activity and PD-L1 presentation. Cell Rep. 2018; 24: 630–41. 138.138. Yang Y, Hsu JM, Sun L, Chan LC, Li CW, Hsu JL, et al. Palmitoylation stabilizes PD-L1 to promote breast tumor growth. Cell Res. 2019; 29: 83–6. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/s41422-018-0124-5&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=30514902&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 139.139. Koster KP, Yoshii A. Depalmitoylation by palmitoyl-protein thioesterase 1 in neuronal health and degeneration. Front Synaptic Neurosci. 2019; 11: 25. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.3389/fnsyn.2019.00025&link_type=DOI) 140.140. Chen M, Andreozzi M, Pockaj B, Barrett MT, Ocal IT, McCullough AE, et al. Development and validation of a novel clinical fluorescence in situ hybridization assay to detect JAK2 and PD-L1 amplification: a fluorescence in situ hybridization assay for JAK2 and PD-L1 amplification. Mod Pathol. 2017; 30: 1516–26. 141.141. Gorinski N, Wojciechowski D, Guseva D, Abdel Galil D, Mueller FE, Wirth A, et al. DHHC7-mediated palmitoylation of the accessory protein barttin critically regulates the functions of ClC-K chloride channels. J Biol Chem. 2020; 295: 5970–83. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamJjIjtzOjU6InJlc2lkIjtzOjExOiIyOTUvMTgvNTk3MCI7czo0OiJhdG9tIjtzOjE4OiIvY2JtLzE3LzMvNTgzLmF0b20iO31zOjg6ImZyYWdtZW50IjtzOjA6IiI7fQ==) 142.142. Runkle KB, Kharbanda A, Stypulkowski E, Cao XJ, Wang W, Garcia BA, et al. Inhibition of DHHC20-mediated EGFR palmitoylation creates a dependence on EGFR signaling. Mol Cell. 2016; 62: 385–96. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.molcel.2016.04.003&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=27153536&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 143.143. Tukachinsky H, Petrov K, Watanabe M, Salic A. Mechanism of inhibition of the tumor suppressor Patched by Sonic Hedgehog. Proc Natl Acad Sci USA. 2016; 113: E5866–E75. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NDoicG5hcyI7czo1OiJyZXNpZCI7czoxMjoiMTEzLzQwL0U1ODY2IjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTcvMy81ODMuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 144.144. Taguchi T, Misaki R. Palmitoylation pilots ras to recycling endosomes. Small GTPases. 2011; 2: 82–4. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.4161/sgtp.2.2.15245&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21776406&link_type=MED&atom=%2Fcbm%2F17%2F3%2F583.atom) 145.145. Siddiqui MA, Singh S, Malhotra P, Chitnis CE. Plasmodium falciparum protein S-palmitoylation is responsive to external signals and plays a regulatory role in microneme secretion in merozoites. ACS Infect Dis. 2020; 6: 379–92.