Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

nature reviews cancer https://doi.org/10.

1038/s41568-023-00611-4

Review article Check for updates

Transfer RNAs as dynamic


and critical regulators of
cancer progression
Alexandra M. Pinzaru & Sohail F. Tavazoie
Abstract Sections

Transfer RNAs (tRNAs) have been historically viewed as non-dynamic Introduction

adaptors that decode the genetic code into proteins. Recent work Variation in tRNA pools and its
has uncovered dynamic regulatory roles for these fascinating relevance to cancer

molecules. Advances in tRNA detection methods have revealed that Genetic and transcriptional
tRNA deregulation in cancer
specific tRNAs can become modulated upon DNA copy number
and chromatin alterations and can also be perturbed by oncogenic Altered tRNA stability
in cancer
signalling and transcriptional regulators in cancer cells or the
Microenvironmental
tumour microenvironment. Such alterations in the levels of specific influences on tRNAs in cancer
tRNAs have been shown to causally impact cancer progression,
Roles of aminoacyl-tRNA
including metastasis. Moreover, sequencing methods have identified synthetases in cancer
tRNA-derived small RNAs that influence various aspects of cancer
Functions of tRNA fragments
progression, such as cell proliferation and invasion, and could serve in cancer
as diagnostic and prognostic biomarkers or putative therapeutic Conclusions and future
targets in various cancers. Finally, there is accumulating evidence, perspectives

including from genetic models, that specific tRNA synthetases — the


enzymes responsible for charging tRNAs with amino acids — can either
promote or suppress tumour formation. In this Review, we provide an
overview of how deregulation of tRNAs influences cancer formation
and progression.

Laboratory of Systems Cancer Biology, The Rockefeller University, New York, NY, USA. e-mail: apinzaru@
rockefeller.edu; sohail.tavazoie@rockefeller.edu

Nature Reviews Cancer


Review article

Introduction that have been identified so far, approximately 80% occur in tRNAs22.
Transfer RNAs (tRNAs) are small non-coding RNAs that are essential These modifications, the most common of which are base methy­
to protein synthesis, acting as the adaptor molecules that translate lations, modulate the affinity of the anticodon–codon interaction, help
the codon information in messenger RNAs (mRNAs) into the amino to maintain the fidelity of the open reading frame, ensure correct tRNA
acid sequences of proteins. First postulated by Francis Crick in his recognition and discrimination by aaRSs, and impact tRNA folding
‘adaptor hypothesis’, tRNAs were discovered in the late 1950s in bio- and stability (reviewed in refs. 5,23–25). In eukaryotes, it has been esti-
chemical studies1–3. Decades of research elucidated the central role that mated that each tRNA molecule contains on average 13 modifications23.
tRNAs have in protein synthesis and provided a foundation for RNA The types, patterns and extent of modifications differ between tRNA
research more broadly. tRNA studies inspired the development of many species; even within the same tRNA isotype, the levels of modifica-
techniques that are now commonly used to study other types of RNA tion of a specific base can vary substantially depending on cellular
and RNA–protein interactions, such as methods to structurally map conditions5,23,26. This generates complexity within tRNA pools, as cells
RNAs and RNA–protein photocrosslinking4. Human cells have two dis- with the same tRNA transcriptional expression patterns can vary in
tinct pools of tRNAs: a cytosolic pool encoded by nuclear tRNA genes, terms of the composition of their functional tRNA pools depending
and a mitochondrial pool encoded by 22 mitochondrial tRNA (mt-tRNA) on their modification status and cellular conditions. An important
genes. True to their postulated evolutionary origins, mt-tRNAs share source of functional variation in tRNAs comes from modifications
features with bacterial tRNAs that deviate from the cytosolic tRNAs, that modulate the decoding ability of tRNAs. For example, inosine,
with peculiarities in their genetic code and secondary structures. Here a deaminated adenosine, can pair with adenosine (A), cytidine (C)
we focus on cytosolic tRNA (hereafter referred to as tRNA) and we refer and uridine (U), and its presence at position 34 in the anticodon of
the reader to recent reviews for detailed discussion on mt-tRNA5–7. the yeast alanine tRNA8 prompted Francis Crick to develop the ‘wobble
The main function of tRNAs is to accept activated amino acids hypothesis’27. Wobble pairing expands the decoding ability of tRNAs,
from aminoacyl-tRNA synthetases (aaRSs) and subsequently transfer allowing all 61 codons corresponding to the standard 20 amino acids
them to a growing peptide chain at the ribosome upon specific codon to be decoded even though certain codons are missing cognate tRNA
recognition during protein translation. This function is achieved due genes, as humans have only 47 tRNA families with distinct anticodons
to the unique structural features of tRNAs. In eukaryotes, tRNAs are according to current in silico predictions28 (Table 1).
short nucleic acid sequences, approximately 73–100 nucleotides, that The human genome encodes approximately 619 predicted tRNA
fold into a ‘cloverleaf’ secondary structure8 and an L-shaped tertiary genes, out of which 429 are of high confidence, meaning they pass all
structure9,10 (Fig. 1a). One arm of the ‘L’ contains the aminoacylation site, the filtering criteria of the current prediction algorithm that distin-
whereas the second arm includes the three-nucleotide anticodon neces- guishes bona fide tRNA genes that function in protein translation from
sary for codon recognition on mRNAs (Fig. 1a). To ensure an adequate tRNA-derived repetitive elements in the human genome28,29 (Table 1).
supply of tRNAs that match the proteomic needs of a cell, tRNAs are tran- tRNAs that carry the same amino acid but have different anticodons
scribed by the specialized RNA polymerase III (Pol III)11 (Fig. 1b). Canoni- are called isoacceptors. Furthermore, an isoacceptor family consists
cally, tRNA genes contain internal promoter sequences (Fig. 1b), except of a variable number of isodecoders, which are tRNAs that share the
for the selenocysteine tRNA gene, which uses extragenic promoter same anticodon, but have sequence differences in the body of the tRNA.
elements, including a TATA box upstream of the transcription start These features contribute to the large diversity of tRNA sequences
site11. Following transcription, the produced precursor tRNA (pre-tRNA) and the complexity of the tRNAome, the regulation of which remains
undergoes extensive processing, including removal of 5′ leader and 3′ an active area of investigation. The number of isodecoders a given
trailer sequences, 3′-CCA addition and common modifications, as well isoacceptor family can have varies greatly. For example, the tRNACysGCA
as quality control to generate a functional mature tRNA12 (Fig. 1c). isoacceptor family comprises 23 different isodecoders, encoded by
For many decades, owing to their essential function in translation, 29 genes, whereas the tRNAHisGTG isoacceptor has only one isodecoder
tRNAs were considered to be static molecular effectors in cells. However, that is encoded by 9 duplicated genes28 (Table 1). This tRNA diversity
data from mass spectrometry, hybridization-based techniques such as is thought to have evolved to match the dynamic proteomic needs of a
microarrays, next-generation sequencing approaches and Pol III occu- complex organism such as a human, and deregulation of this diversity
pancy analysis have uncovered a dynamic pattern of tRNA expression is implicated in disease. For example, tRNA levels have been found to
that is dependent on tissue type, developmental stage and prolifera- be altered in cancer cells compared with normal cells, probably as
tive state, and have revealed that tRNAs can have regulatory roles and a means to optimize protein production to sustain the growth and
respond to environmental cues to shape gene expression13–21. The grow- survival of cancer cells18,30. Moreover, as detailed below, there is evi-
ing appreciation of tRNAs as important post-transcriptional regulators dence of specific tRNAs regulating cell proliferation31 and metastatic
of gene expression, aided by an expanding toolkit to study and manipu- progression21,32, functionally implicating tRNAs in various steps of
late tRNA, has sparked great interest in the past 15 years to uncover cancer initiation and progression (Fig. 2).
the previously unexplored roles of tRNA in cancer. In this Review, we
describe loss-of-function, gain-of-function and clinical association Functional consequences of altered tRNA pools in cancer
studies that provide evidence for canonical and non-canonical roles of In the late 1960s, chromatography-based techniques began hint-
tRNAs in cancer. We also detail recent findings suggesting that aaRSs ing at differences in the levels of tRNAs between tumour cells and
can act as tumour suppressors or promoters of cancer progression. untransformed cells33–35. However, it was not until 2009 that a reliable
genome-wide assessment of tRNA expression in cancer was reported30.
Variation in tRNA pools and its relevance to cancer Using microarrays to assess the levels of 40 nuclear-encoded isoac-
A distinctive and fascinating feature of tRNAs is their extensive incor- ceptors in normal and cancerous breast epithelial cells as well as
poration of modified nucleosides, which are crucial for tRNA stability patient-derived tumours, the authors found a consistent trend of over-
and function. Of the approximately 150 ribonucleoside modifications expression of tRNAs in breast cancer cell lines and tissues relative to

Nature Reviews Cancer


Review article

normal controls. In addition, a selective upregulation of certain tRNA genes30. A later study from a different group used custom-made micro-
isoacceptors was detected, including tRNAArgCCU, tRNAThrCGU, tRNALeuUAA, arrays to profile in parallel the levels of 206 tRNA genes and approxi-
tRNATyrGUA, tRNASerGCU and tRNAArgUCU, which the authors speculated mately 7,000 mRNAs from an extensive set of primary tumours and
may be necessary for the efficient translation of tumorigenesis-related cancer cell lines, including bladder, colon and prostate cancers, as

a 3′
A
C
C
5′ N73

Acceptor
stem TψC-
TψC- loop
D-loop stem
D-stem

Variable
Anticodon region
stem

Anticodon
loop

Anticodon 34 35 36

b c RNase P RNase Z

3′
5′ Trailer
Pol III TFIIIC Leader All tRNAs
• 5′ Leader and 3′
5′ TBP BRF1 trailer removal
TSS
3′ BDP1 • Common
A box B box TTTTT modifications
tRNA • 3′-CCA addition

Specific tRNAs
• Intron splicing
Intron • Specific
modifications
• Aminoacylation

Fig. 1 | Transfer RNA structure and biogenesis. a, Schematic on the left protein (TBP), transcription factor TFIIIB component B″ homologue (BDP1) and
shows the classical transfer RNA (tRNA) cloverleaf secondary structure with the transcription factor IIIB 90 kDa subunit (BRF1). DNA-bound TFIIIB and TFIIIC
acceptor stem, the variable region and the three arms (each consisting of a stem recruit the specialized RNA polymerase III (Pol III)11 to form the pre-initiation
and a loop region), including the anticodon arm, the dihydrouridine-containing complex (PIC) at the tRNA genes. Structural rearrangements of the PIC allow
(D)-arm, and the thymidine, pseudouridine and cytidine-containing (TψC)- transcription initiation and promoter escape, followed by transcription
arm. The anticodon loop includes in its centre, at positions 34–36 (5′ → 3′), the elongation until a termination signal, a short stretch of at least five thymidine
anticodon necessary for codon recognition on mRNAs. The 3′ end of the acceptor nucleotides in human cells on the non-template strand, is encountered. Upon
stem ends in an invariant CCA triplet that contains the aminoacylation site on transcription termination, Pol III is recycled to rapidly reinitiate transcription
the final adenosine167 with an upstream discriminator base (N73), which is used of the same gene, resulting in high yields of tRNA transcripts11. c, Processing of
as a recognition element by aminoacyl-tRNA synthetases168. The acceptor stem, the precursor tRNA (pre-tRNA) involves cleavage of the extra 5′ leader and 3′
anticodon arm and the TψC-arm have conserved sizes, whereas the D-arm and the trailer sequences by the ribonucleoprotein complex ribonuclease P (RNase P)
variable loop sizes can differ, accounting for the variation in tRNA length167,169 (not and RNase Z12 (for which two forms in human cells, ELAC1 and ELAC2, have been
shown). On the right is shown the tertiary structure of a tRNA based on the crystal identified173), respectively. In humans and most eukaryotes, tRNA nucleotidyl
structure of yeast phenylalanine tRNA170 (Protein Data Bank ID: 1EHZ) (the 3D transferase 1 (TRNT1) adds the invariant 3′-CCA triplet after the discriminator
rendition was created using the UCSF ChimeraX 1.4 software). Base stacking and base (not shown). A subset of tRNAs contain intronic regions, which are excised
interactions between conserved or semi-conserved residues, mainly between the by splicing enzyme complexes, before the tRNAs undergo final modifications
D-loops and TψC-loops, lead to the 3D L-shape of the tRNA. The TψC-stem stacks that allow them to assume a functional tertiary structure12. Maturation of tRNAs
onto the acceptor stem to form one arm of the ‘L’, whereas the anticodon stem– includes incorporation of common modifications, such as pseudouridine and
loop stacks onto the D-stem to form the second arm of the ‘L’. The colour scheme dihydrouridine, whereas certain tRNAs incorporate specific modifications
matches the one used in the secondary structure. b, tRNA genes contain internal such as anticodon modifications inosine and queuosine5,12. Finally, each mature
promoter sequences called the A box and B box171,172 that are recognized by the tRNA will undergo aminoacylation, the process in which the tRNA is covalently
multisubunit general transcription factor IIIC (TFIIIC), which then recruits bound by a dedicated aminoacyl-tRNA synthetase to a specific amino acid
the TFIIIB transcription factor, a trimeric complex composed of TATA-box binding corresponding to its anticodon108. TSS, transcription start site.

Nature Reviews Cancer


Review article

Table 1 | High-confidence tRNA genes and corresponding isodecoders (hg38)a

Isoacceptor Anticodon Number of tRNA Number of distinct Isoacceptor Anticodon Number of tRNA Number of distinct
genes isodecoders genes isodecoders

Ala AGC 26 17 Tyr GTA 13 9


CGC 4 4 Val AAC 9 5
TGC 8 7 CAC 13 7
Arg ACG 7 2 TAC 5 4
CCG 4 2 iMet, initiator Met; tRNA, transfer RNA. aData curated from GtRNAdb28.

TCG 6 6
well as glioblastomas and diffuse large B cell lymphomas (DLBCLs)18.
CCT 5 5
The authors compared tRNA pools in cancer samples relative to nor-
TCT 6 5 mal tissue controls and found that patients within the same cancer
Asn GTT 25 16 type tended to have similar tRNA pools, but that similarity decreased
when comparing different cancer types to each other. Moreover, by
Asp GTC 13 3
experimentally altering cell proliferation and differentiation in vari-
Cys GCA 29 23
ous cellular systems, the authors identified ‘differentiation tRNAs’ and
Gln CTG 13 7 ‘proliferation tRNAs’, the tRNA availability of which was correlated with
TTG 6 4 the distinct codon usage of proliferation genes versus differentiation
genes. In addition, the authors used their DLBCL data to detail specific
Glu CTC 8 2
tRNA changes and reported that, relative to normal B cells, DLBCL
TTC 8 5 samples exhibited consistent changes in their tRNA pools, with several
Gly GCC 14 5 tRNAs showing over tenfold enrichment, such as tRNALysCUU, whereas
CCC 5 3
other tRNAs exhibited more than tenfold depletion, such as tRNATyrGUA.
Finally, across different cancer types, the authors also revealed correla-
TCC 9 4
tions between the levels of upregulated tRNAs in cancer samples with
His GTG 9 1 the codon usage of transcripts induced in cancer18.
Ile AAT 15 10 The early explorations of the tRNA profiles in cancer cells relative
to normal cells have provided important clues to how tRNA mole­
GAT 3 1
cules might affect cancer progression. However, loss-of-function
TAT 5 3 and gain-of-function analyses are necessary to establish causative
iMet/Met CAT 9/11 2/7 roles for tRNA changes during tumorigenesis (Box 1). One of the first
Leu AAG 9 4 tRNAs to be experimentally modulated was the initiator methionine
(tRNAiMet), which has an essential role during translation initiation —
CAG 9 2
the rate-limiting step during protein synthesis36 — and is upregulated
TAG 3 3 in breast cancer samples30. It has been shown that the overexpression
CAA 6 5 of mature tRNAiMet in untransformed human breast epithelial cells
increased metabolic activity and proliferation in those cells31. The
TAA 4 4
mechanism underlying this observation was not investigated, but
Lys CTT 15 10 the authors reported that overexpressing tRNAiMet increased the levels
TTT 12 8 of other tRNAs, emphasizing the complexity of tRNA regulation31. In a
Phe GAA 10 5 different study, overexpression of tRNAiMet in a highly metastatic human
melanoma cell line did not affect primary tumour growth but increased
Pro AGG 9 2
migration and invasion of cancer cells37. These studies have revealed
CGG 4 2 that overexpression of tRNAiMet promotes tumorigenic phenotypes
TGG 7 3 consistent with enhancement of global translation eliciting malignant
SelCys TCA 1 1
phenotypes38,39. Furthermore, these results also suggest that the spe-
cific effects of tRNA overexpression on cancer progression may be con-
Ser AGA 9 4
text dependent, causing different phenotypes depending on cell type.
CGA 4 4 In addition to cancer cells, increased levels of tRNAiMet have also been
TGA 4 4 reported in cancer-associated fibroblasts, and the potential effects
of this dysregulation in the tumour stroma have been investigated
GCT 8 6
by introducing two copies of the gene encoding tRNAiMet upstream of
Thr AGT 9 6 exon 1 of the Hprt gene in a mouse model (2 + tRNAiMet mice). The authors
CGT 5 5 found that syngeneic tumours injected into 2 + tRNAiMet mice grew
TGT 6 6
faster and exhibited greater vascularization than tumours injected
into littermate controls. These tumour phenotypes were linked to
Trp CCA 7 5
preferential altering of the secretome, as mouse embryonic fibroblasts

Nature Reviews Cancer


Review article

derived from the 2 + tRNAiMet mice exhibited increased production and breast tumours compared with non-metastatic primary tumours,
secretion of collagen proteins40. These studies have suggested that revealing that their expression correlated with human breast cancer
tRNAiMet can promote specific tumorigenic phenotypes in a cell-type- metastatic potential. Integrating data from ribosome profiling, prot-
dependent manner (Fig. 2) and raise the question of how tRNAiMet levels eomics and whole-genome measurements of mRNA stability revealed
are regulated in cancer. As detailed below, such regulation may be both that tRNAGluUUC and tRNAArgCCG overexpression specifically promoted
transcriptional41 and post-transcriptional42. the translation of mRNAs enriched in codons cognate to the overex-
To probe the question of whether specific tRNA alterations could pressed tRNAs. Two such genes, exosome component 2 (EXOSC2) and
change distinct downstream translational programs during cancer GRIP1-associated protein 1 (GRIPAP1), were identified as downstream
progression, a hybridization-based, high-throughput tRNA profiling mediators of the pro-metastatic effects of tRNAGluUUC21. Together, this
method was developed to investigate tRNA dynamics during cancer study demonstrated that upregulation of specific tRNAs can not only
progression21. Unbiased quantification of isoacceptor levels revealed regulate specific pathways but also causally drive invasive and meta-
that tRNAGluUUC and tRNAArgCCG levels were consistently upregulated static phenotypes through enhanced translation of a specific down-
in highly metastatic breast cancer cells compared with their isogenic stream gene network comprising genes enriched in the codons cognate
parental lines21. Depletion and overexpression studies found that these to the upregulated tRNAs (Fig. 2).
two tRNAs promoted the invasiveness and enhanced the metastatic Despite accumulating evidence that tRNAs can regulate cancer
capacity of breast cancer cells. Moreover, tRNAGluUUC and tRNAArgCCG progression, systematic studies that detail changes in tRNA levels dur-
levels were observed to be upregulated in metastatic primary human ing different stages of cancer evolution across different cancer types

Leu
Normal cell

CAG
1 Transformation
Arg

UCU

Cancerous cell
iMet (poorly metastatic) Glu Arg Ile Ile

CAT UUC CCG GAU UAU


2 Tumour Metastasis 3
growth Ribosome
CGG rich
mRNA AUC rich
GAR rich
AUA rich
Metastatic
progression
Metastatic
↑ Translation
progression

? Specific
↑ Proliferation proteins

Fig. 2 | Roles of mature transfer RNAs in oncogenic transformation and including increased proliferation31 by unknown mechanisms (indicated by ‘?’).
cancer progression. Model showing how mature transfer RNA (tRNA) levels In mouse cancer-associated fibroblasts, tRNAiMet upregulation leads to
can influence tumour initiation and progression. Shown are examples of increased production of specific proteins, which increases tumour growth40.
tRNAs reported to be involved at each stage during cancer progression. During metastatic progression (3), altered levels of specific tRNAs promote the
Oncogenic transformation (1) can be suppressed in oncogene-expressing breast translation of the mRNAs enriched in the codons cognate to the upregulated
epithelial cells by tRNALeuCAG (ref. 103), or promoted in simian virus 40 (SV40) tRNAs21,32. These can include enhanced translation of GAR-rich (GAA and
immortalized mouse embryonic fibroblasts or in NrasG12D/+ haematopoietic stem GAG codons) or CGG-rich mRNAs in response to upregulation of tRNAGluUUC
or progenitor cells by ectopic expression of tRNAArgUCU (ref. 81). Tumour growth and tRNAArgCCG21 or preferential translation of AUA-rich over AUC-rich
(2) has been shown to be influenced by various tRNAs18,30,31,40,81–85. Increased mRNAs in response to changes in the ratio of tRNAIleGAU to tRNAIleUAU (ref. 32).
tRNAiMet levels can promote mRNA translation174 with cell-type-specific effects, mRNA, messenger RNA.

Nature Reviews Cancer


Review article

are still lacking. One limitation of such studies still lies in achieving quantification of tRNAs is still not possible, recent strides have been
accurate tRNA quantification due to the intricate modification pat- made in circumventing some of the hurdles imposed by tRNA modi-
terns on tRNAs that block reverse transcription, their rigid secondary fications and provide improved relative tRNA quantification at the
structures as well as high sequence similarity. Even though absolute isodecoder level16,43–46 (Box 1 and reviewed in ref. 23).

Box 1

Challenges in measuring and experimentally modulating tRNA


levels
Historically, one of the major hurdles in studying transfer RNAs gene knockouts for all seven murine isodecoder genes encoding
(tRNAs) in cancer has been obtaining high-resolution, accurate tRNAPheGAA (ref. 180). In line with the expected redundancy of related
quantification of tRNA levels23. Next-generation sequencing offers isodecoders, the single deletions of six out of the seven isodecoders
accessible, relatively inexpensive platforms for quantifying tRNAs, did not cause overt phenotypes. Only the single deletion of
but library preparation for next-generation sequencing relies on tRNA-Phe-GAA-1-1 caused subtle defects in growth and neurological
reverse transcription of the tRNA molecules. The stable folding development180. Knockout of tRNA-Phe-GAA-1-1 worsened the
and modification of tRNAs cause the reverse transcriptase to stop embryonic lethality phenotype observed upon simultaneous
before reaching the end of the molecules, resulting in many short, deletion of multiple tRNA-Phe-GAA isodecoders, suggesting that
incomplete reads, which preclude accurate quantification of mature not all related tRNA isodecoder genes are functionally redundant.
tRNA levels. This problem is compounded by the high sequence Interestingly, the reduction of mature tRNAPheGAA in the single
similarity of tRNA genes, which can differ by only a few nucleotides deletion mutants was tissue dependent, with tissues such as the
for different isoacceptors28, meaning the short reads can map to kidney showing no depletion, whereas others, such as the liver, had
multiple tRNA genes. For these reasons, we caution against the more than 70% reduction in mature tRNAPheGAA levels180. This data,
inference of mature tRNA levels from small-RNA sequencing data together with previous reports that different tissues have different
not geared towards full-length tRNA expression identification, such tRNA expression patterns15,16, highlight the need to systematically
as small-RNA sequencing datasets from The Cancer Genome Atlas, investigate the contribution of each tRNA isodecoder gene to the
which are specifically aimed at identifying microRNAs and have read pool of functional tRNAs for a tissue of interest. Conversely, following
lengths of only approximately 30 nucleotides175. identification of decreased mature tRNA levels in cancer cells, it is
Different strategies for overcoming the reverse transcription necessary to determine the exact tRNA genes that are downregulated
blocks have emerged, including using higher processivity reverse to properly model them in loss-of-function studies.
transcriptases that can read through some of the modifications Gain-of-function experiments are also complicated, as it can be
and/or by enzymatically removing certain modifications altogether, challenging to achieve detectable levels of tRNA overexpression31,149.
where possible23,176,177. Interestingly, next-generation sequencing- For example, introducing 10 or 20 copies of tRNA-Gly-GCC in a
based approaches have also been developed to detect and quantify Drosophila model with glycyl-tRNA synthetase 1 (GARS1) mutations
a subset of tRNA modifications23,44. Some tRNA profiling methods use only increased mature tRNAGlyGCC levels by approximately 13% and
hybridization approaches to avoid the reverse transcription step approximately 25%, respectively, whereas 12 copies of tRNA-Gly-TTC
altogether15,21, but they lack the sensitivity to distinguish between achieved up to 75% overexpression of tRNAGlyUUC (ref. 149). In mouse
tRNA genes that differ by less than eight nucleotides15. An approach models, two copies of a transgene containing a genomic locus with
that holds high promise to retrieve both the full tRNA sequences two tRNA-Gly-GCC genes was not enough to increase mature tRNA
and their modification status (but not necessarily the identity of the levels, but expressing approximately 27 copies of that transgene
modification) is direct sequencing using nanopore sequencing achieved 90–150% increase in mature tRNAGlyGCC levels, depending
technology46,178,179. This is a developing technology that has very on tissue type149. These results suggest that the exact number of gene
recently been shown in yeast to have potential for tRNA quantifica­ copies necessary to sufficiently increase mature tRNA levels vary
tion46 and — when optimized for human tRNA — might streamline for each tRNA isoacceptor and probably depend on the studied cell
detection of tRNA levels in cancer samples. type. In support of that idea, transgenic mice carrying two additional
To show causal roles for tRNAs in cancer, it is critical to generate copies of the gene encoding tRNAiMet had approximately 1.3–1.5-fold
the appropriate loss-of-function or gain-of-function models of higher levels of tRNAiMet in various tissues40. To note, the integration
disease. However, tRNAs present unique challenges in achieving of the two copies of the gene encoding tRNAiMet in the Hprt locus in
robust modulation of their levels. As opposed to most coding genes, an open chromatin region overlapping with the Hprt promoter region
which have just two alleles, all tRNAs except tRNASec are encoded upstream of exon 1 might have favoured the transcription of the tRNA
by multigene families (Table 1), challenging attempts to create tRNA genes40. The use of an upstream, strong RNA polymerase III promoter
isotype knockouts, especially when generating mouse models, as one such as the U6 promoter has been successfully implemented to
may need to edit several genomic loci at once to achieve sufficient drive higher tRNA gene expression than just by overexpressing the
depletion of the tRNA of interest. This was recently exemplified in endogenous tRNA genes181 and may be combined with multiple tRNA
a study using CRISPR–Cas9 to generate single and combinatorial copies to optimize overexpression systems150.

Nature Reviews Cancer


Review article

1 Changes in cancer 2 Effects on tRNA 3 Effects on translation

Mutations: ACTAGAAAT > ACTAGCAAT Altered decoding Amino acid misincorporation


Copy number changes: Polypeptide
Ala Ser Ser Ala
Genetic changes

chain
Ref tRNA-1
Amplifications Ser
tRNA-1 tRNA-1 mRNA
Ala
AGC AGA AGC Ribosome
codon
Ref tRNA-2
Deletions Altered tRNA pools Codon-biased translation
Deletion

DNA Me
MYC
methylation
Transcriptional changes

Pol III TFIIIC


5′ SOX4 TBP BRF1
3′ BDP1 A box B box TTTTT

tRNA

p53 pRb

tRNA
Modification changes degradation
Post-transcriptional
deregulation

Amino acid charging levels

tRNA mischarging Amino acid misincorporation


Trp Phe
Trp
Phe
Amino acid deprivation WARS1
Environmental

CCA CCA Trp codon


stressors

Hypoxia O2
tRNA Non-canonical functions
O2 O2 fragmentation (e.g., regulation of mRNA stability)

Fig. 3 | Regulation of transfer RNA expression in cancer. Diagram summarizing BRF1, transcription factor IIIB 90 kDa subunit; Me, methylation; Pol III, RNA
the processes (1) that can lead to transfer RNA (tRNA) deregulation in cancer polymerase III; Ref, reference genome; SOX4, SRY-box transcription factor 4;
(2), and affect tRNA decoding capabilities, abundance, stability or charging TBP, TATA-box binding protein; TFIIIC, general transcription factor IIIC;
by aminoacyl-tRNA synthetases resulting in altered messenger RNA (mRNA) TSS, transcription start site; WARS1, tryptophanyl-tRNA synthetase 1.
translation (3). BDP1, transcription factor TFIIIB component B″ homologue;

Genetic and transcriptional tRNA deregulation affect the folding, stability and decoding affinity of the tRNA. Indeed,
in cancer there are instances in which single tRNA mutations have been found
Copy number variation and mutations in tRNA genes to contribute to disease phenotypes. For example, a hypomorphic
The observation that tRNAs have prominent roles in tumour pro- single-nucleotide polymorphism in one tRNA-Arg-TCT gene caused
gression raises the question of how tRNA levels are altered in cancer ribosome stalling and neurodegeneration in mice when combined
(Fig. 3). Somatic copy number variation has been shown to be one with loss of the GTP-binding protein 2 (GTPBP2) ribosome recycling
mechanism responsible for tRNA upregulation during cancer evolu- factor52. In humans, a loss-of-function homozygous single-nucleotide
tion to a highly metastatic state21 (Fig. 3). As germline copy number polymorphism in the tRNA-SeC-TCA gene preferentially impaired the
changes have been reported for certain tRNA genes in humans47,48, production of stress-related selenoproteins, presenting with various
it is tempting to speculate that germline copy number variations of clinical symptoms in a patient53. In proof-of-concept experiments
tRNA genes may influence cancer progression, although their func- relevant to cancer, exogenous expression of a tRNASer with a single
tional implications remain to be investigated. Apart from copy num- point mutation that incorporates serine at specific alanine codons
ber variations, tRNA genes exhibit large sequence variation in both (tRNASer(Ala)) has been shown to increase the tumorigenic potential of
their tRNA bodies49,50 and their immediate flanking regions (approxi- NIH-3T3 mouse embryonic fibroblasts54. Compared with control cells,
mately 20 bp upstream and approximately 10 bp downstream from the NIH-3T3 cells expressing tRNASer(Ala) had significantly accelerated
the mature tRNA)51. Polymorphisms in the body of the tRNA could tumour growth in xenografts assays in nude mice. To note, the authors
impact the efficiency and fidelity of protein translation, as they can observed spontaneous transformation of their control NIH-3T3 cells,

Nature Reviews Cancer


Review article

Glossary genes were occupied across all tested cell lines, these experiments also
identified cell-type-specific patterns of Pol III occupancy in human
cells19,20. Evidence in support of transcriptional control of specific tRNA
Argonaute 2 (AGO2) Ribosome profiling genes during cancer progression was observed for tRNAIleUAU, as highly
cleavage assay A technique that provides a snapshot metastatic breast cancer cells exhibited increased Pol III occupancy
An in vitro assay that uses recombinant of all ribosome-bound messenger RNAs at tRNA-Ile-TAT genes relative to poorly metastatic cells, which resulted
AGO2 to show specific cleavage of in cells with codon-level resolution. in increased tRNAIleUAU expression32. Interestingly, the upregulation
target oligonucleotides. of tRNAIleUAU was accompanied by repression of the related isoacceptor
RNA polymerase III tRNAIleGAU during breast cancer progression32. Mimicking the divergent
Epistatic interactions (Pol III). An RNA polymerase complex modulation of the two related tRNAIle isoacceptors experimentally
Interactions between two or more that translates short, non-coding RNAs, promoted metastatic colonization in breast cancer xenograft models.
genes that influence a particular including transfer RNAs, 5S ribosomal Proteomics and ribosome profiling revealed that the deregulation
phenotype. RNA, U6 small nuclear RNA, vault RNA of these two tRNAIle isoacceptors resulted in increased translation of
and other small RNAs. growth-promoting mRNAs enriched in the AUA codon, which is cognate
Lineage-negative to the upregulated tRNAIleUAU (Fig. 2). Of note, the dually modulated cells
haematopoietic stem tRNA misacylation had a distinct proliferation advantage relative to control cells under
and progenitor cells The charging of a transfer RNA hypoxia and oxidative stress: two key features of the metastatic tumour
Immature blood cells that lack with the wrong amino acid, not the environment. Moreover, repression of tRNAIleGAU and upregulation of
surface protein markers found on amino acid cognate to the anticodon tRNAIleUAU have been observed in human breast tumours that had higher
mature blood cells. of the transfer RNA. rates of metastatic relapse32.
Several tumour suppressor and pro-tumorigenic transcription
Rare codons Wobble pairing factors have been reported to exert opposing effects on Pol III activity,
Codons found with the lowest Non-Watson–Crick base pairing of the although their effects on global tRNA pools versus specific isodecoder
frequency in genes compared with first anticodon nucleoside: inosine can genes have not been analysed in detail57,58. Of note, the tumour suppres-
synonymous codons in the same group. pair with adenosine, uridine or cytidine; sors p53 and pRb were both reported to bind to TFIIIIB and prevent
guanosine can pair with cytidine or recruitment of Pol III to its target genes59–61, whereas binding of the
Ribosome uridine; uridine can pair with adenosine MYC oncoprotein to TFIIIB stimulates the Pol III promoter62,63 (Fig. 3).
An essential ribonucleoprotein complex or guanosine. Moreover, Pol III activity is also stimulated by the mechanistic target
that performs protein synthesis in cells. of rapamycin (mTOR) and Ras oncogenic pathways64. Despite ChIP–
seq data showing overlap of several other transcription factors with
Pol III sites19,65, functional evidence of a transcription factor directly
although with much later latency than the tRNASer(Ala)-expressing cells binding upstream of a subset of tRNA genes has only been recently
(Fig. 3). Interestingly, the misincorporation of serine at alanine codons provided41. The authors found that overexpression of the transcrip-
did not impact cell viability in vitro but did provide NIH-3T3 cells with tion factor SRY-box transcription factor 4 (SOX4) in glioblastoma cells
an advantage during tumour growth in vivo54. The contribution of caused cell cycle exit41, in line with previous work showing that SOX4
particular tRNA variants to cancer has not been formally investigated expression can contribute to the reprogramming of glioblastoma cells
thus far, especially as the effects from mutating one tRNA gene are to a more terminally differentiated state66,67. Surprisingly, ChIP–seq
expected to be alleviated by the redundant copies of the tRNA genes uncovered SOX4 binding to over 200 tRNA gene loci in glioblastoma
in the same isoacceptor family. Nonetheless, as the aforementioned cells and mechanistic analysis has found that SOX4 binding reduced
studies suggest, the effects of tRNA variants may become apparent the recruitment of TFIIIB and Pol III at tRNA promoters and repressed
via epistatic interactions with specific mutated cancer genomes and/or target tRNA genes. Using functional assays, the authors showed that
under stress conditions, such as that of the tumour microenvironment SOX4 bound to a motif upstream of the tRNA-iMet-CAT-1 isodecoder
or during metastatic progression. family and repressed transcription. However, SOX4 did not alter the
Mutations in the internal A and B boxes, as well as in the immedi- expression of the related tRNA-iMet-CAT-2 isodecoder, and ectopic
ate flanking regions of the tRNAs, may also alter the transcriptional overexpression of tRNA-iMet-CAT-2 partially rescued the total tRNAiMet
output of tRNA genes. It was already appreciated in the 1980s that levels and the proliferation defect induced by SOX4 overexpression41.
transcription of tRNA genes can be modulated by 5′ upstream flanking It is unclear whether all the putative tRNA genes targeted have similar
regions outside of the conserved internal promoter elements in various SOX4-binding motifs and whether they are all responsive to SOX4 bind-
organisms, including humans55,56. However, to our knowledge, recur- ing. Nevertheless, this work suggests that specific transcription factors
rent features or functional motifs in the flanking regions of tRNA genes may directly regulate particular tRNA genes during cancer progression
have not been comprehensively and systematically investigated so far. (Fig. 3). Interestingly, SOX4 has been shown to have opposing roles in
cancer progression, acting as a tumour suppressor in certain tumour
Altered Pol III activity deregulates tRNAs in cancer types while having pro-oncogenic and pro-metastatic roles in other
The transcriptional control of tRNA genes depends on Pol III, with Pol III cancer types68,69, which raises the question of whether the dual role
occupancy indicative of active tRNA expression. In contrast to yeast, of SOX4 in tumorigenesis is, at least in part, mediated via differential
where virtually all tRNA genes are expressed, chromatin immunopre- modulation of tRNA expression in different cell types.
cipitation coupled with sequencing (ChIP–seq) experiments have found In addition to transcription factors, Pol III activity at tRNA genes is
that Pol III only occupies approximately 50% of all predicted human also influenced by chromatin accessibility19,20,58. DNA methylation has
tRNA genes. Interestingly, even though a set of approximately 120 tRNA been found to repress tRNA gene transcription in vitro70 and inhibit

Nature Reviews Cancer


Review article

TFIIIC binding to DNA71. Interestingly, DNA methylation has been sug- Mettl1-knockout mouse model revealed that, consistent with the xeno-
gested to have a role in regulating the levels of specific tRNAs in cancer72. graft assays, loss of Mettl1 inhibited HCC85, ESCC82 and HNSCC83 tumo-
Mining DNA methylation data from The Cancer Genome Atlas revealed rigenesis. Mechanistically, depletion of either METTL1 or WDR4 caused
that 60 of the 71 tRNA genes with at least one CpG island covered by hypomodification of specific tRNAs and reduced the overall levels of a
the methylation array that was used exhibited statistically significant subset of m7G-modified tRNAs in glioblastoma cells81 and ICC84, ESCC82,
differences between normal and tumour samples. In addition, 86 cases HCC85 and HNSCC83 cell lines. Moreover, altering METTL1–WDR4
were reported for which the methylation status of a tRNA gene was levels changed the expression of genes related to cell cycle and acti-
associated with significant differences in overall patient survival, with vated oncogenic pathways81–85. Some of these studies81,82,84,85 have
certain tRNA genes showing an effect in more than one cancer type. provided some analyses to show that the METTL1–WDR4 effects on
Functional assays have revealed that the expression of two isodecoder tumour growth are mediated, at least partially, via their downstream
genes, tRNA-Arg-TCT-4-1 and tRNA-Ile-AAT-8-1, was indeed regulated tRNA targets. For example, overexpressing tRNAArgUCU recapitulated
by DNA methylation. Moreover, unmethylated tRNA-Arg-TCT-4-1 has the pro-tumorigenic effects and proteome changes induced by
been shown to be associated with significantly worse overall survival METTL1–WDR4 ectopic expression81. Interestingly, overexpression
outcome of patients with kidney papillary cell carcinoma and uterine of tRNAArgUCU in simian virus 40 T antigen-immortalized mouse embry-
corpus endometrial carcinoma72. These findings suggest that the methy­ onic fibroblasts increased their anchorage-independent growth in
lation of tRNA genes may be an important mechanism for controlling soft agar compared with control cells. Moreover, ectopic expression
specific tRNA gene expression in cancer progression (Fig. 3). of tRNAArgUCU in primary non-leukaemic NrasG12D/+ lineage-negative
haematopoietic stem and progenitor cells promoted colony formation
Altered tRNA stability in cancer during serial replating assays relative to controls, providing in vitro evi-
tRNA levels can also be regulated post-transcriptionally through dence for a pro-tumorigenic role for this isoacceptor81 (Fig. 2). However,
changes in tRNA processing and stability. Mutations in tRNA processing in the absence of exhaustive epistatic interaction analyses between
factors associated with alterations in tRNA stability have been observed METTL1–WDR4 and its target tRNA genes in each studied cancer type,
in several non-malignant human disorders7; however, such mutations it is challenging to draw conclusions about which tRNAs mediate the
have only rarely been implicated in cancer progression, perhaps complex effects of the methyltransferase complex on oncogenesis.
because they would probably result in global reduction of tRNA levels, Moreover, METTL1, as well as most tRNA-modifying enzymes, can act
impaired protein synthesis and growth inhibition. Thus far, only muta- on other RNA species, such as mRNAs and microRNAs (miRNAs), apart
tions in the gene elaC ribonuclease Z 2 (ELAC2), which encodes the long from tRNAs24,86,87. It would be important to disentangle the potential
form of RNase Z in eukaryotes, have been associated with cancer73 and effects of METTL1 depletion on tRNAs versus other RNA species in
may be low-penetrance susceptibility markers of prostate cancer74. future studies, as well as to biochemically assess whether tRNAs lacking
Nucleobase modifications are also an important determinant of m7G are subjected to increased degradation in human cells. Neverthe-
tRNA stability. Modifications outside the anticodon arm are linked to less, these studies provide support for exonuclease-mediated decay
stabilizing tRNA structure and inhibiting tRNA degradation, whereas pathways regulating tRNA levels in cancer (Fig. 3) and motivate the
modifications in the anticodon arm generally regulate tRNA decoding characterization of the molecular details of such pathways.
function24. The degradation of hypomodified tRNAs has mostly been Finally, experimental modulation of tRNA levels using small
studied in Saccharomyces cerevisiae, in which two exonucleolytic path- hairpin RNAs suggested that RNAi pathways could regulate tRNA
ways that mediate the breakdown of hypomodified tRNAs have been expression21,88. A recent study has provided, to our knowledge, the
described. The Trf4–Air2–Mtr4p polyadenylation (TRAMP) complex, first evidence of naturally occurring miRNA-mediated regulation of
the members of which are conserved in humans75, degrades tRNAiMet tRNAs42. The authors found that unmodified pre-tRNAiMet is a target
that lacks N1-methyladenosine at position 58 in a 3′→5′ direction76. of the tumour suppressor miR-34a using in vitro pulldowns and the
Hypomodified and unstable tRNAs are also degraded 5′→3′ by a dedi- Argonaute 2 (AGO2) cleavage assay. The authors provided further
cated, rapid tRNA decay pathway in different yeasts77–79, with circum- in vivo evidence of RNAi-mediated control of tRNAiMet levels by modulat-
stantial evidence of it being conserved across eukaryotes, including in ing miR-34a levels and by using small hairpin RNAs and small interfering
humans80. One of the tRNA body modifications necessary to prevent RNAs targeting tRNAiMet in breast cancer cells. Limited epistasis experi-
tRNA degradation in S. cerevisiae is N7-methylguanosine (m7G) at posi- ments have suggested that the miR-34a cellular phenotypes were
tion 46 (ref. 78). Interestingly, the RNA m7G methyltransferase complex, partially mediated through modulation of tRNAiMet levels42.
comprising methyltransferase-like 1 (METTL1) and WD repeat domain 4
(WDR4), is overexpressed in various cancers81–85. Overexpression Microenvironmental influences on tRNAs in cancer
of wild-type but not catalytically inactive METTL1 promoted malig- The excessive proliferation of cancer cells causes the centre of solid
nant phenotypes in in vitro assays and increased m7G modification tumours to become deprived of nutrients and oxygen. This exposes
on a subset of tRNAs in various cell lines, including mouse embry- cancer cells to environmental stresses, which have been shown to
onic fibroblasts81, intrahepatic cholangiocarcinoma (ICC)84, hepato- regulate tRNAs in yeast14. Some of the most prominent stresses that
cellular carcinoma (HCC)85, oesophageal squamous cell carcinoma cancer cells encounter are hypoxia and amino acid deprivation39.
(ESCC)82, and head and neck squamous cell carcinoma (HNSCC)83 Hypoxia can increase reactive oxygen species production89,90 and thus
cell lines. A conditional Mettl1 knock-in mouse showed in vivo that oxidative stress. The effects of reactive oxygen species on tRNAs are
increased METTL1 expression promoted ESCC82, HCC85 and HNSCC83 multifaceted91. In brief, oxidative stress can change the modification
tumorigenesis. Conversely, depletion of METTL1 or WDR4 inhibited patterns of tRNAs, affecting their decoding abilities91–93. Increased
tumour growth in mouse xenograft assays using various cancer lines, levels of reactive oxygen species also affect tRNAs in non-canonical
including acute myeloid leukaemia, glioblastoma, liposarcoma81, ways through increased mischarging of non-Met-tRNAs with Met94,
ICC cells84, HCC cells85, ESCC cells82 and HNSCC cells83. A conditional selective retrograde transport of tRNAs to the nucleus95 and tRNA

Nature Reviews Cancer


Review article

fragmentation96–99. These tRNA-related response pathways are believed developmental disorders104. Although aaRSs have functions not related
to help protect the cells from the deleterious effects of oxidative stress to protein synthesis that have been implicated in cancer (reviewed in
and raise the question of whether they are altered in cancer cells as a refs. 105–108), here we detail the involvement of aaRSs in cancer based
means for adapting to hypoxic conditions. on their canonical protein synthesis functions.
Amino acid deprivation can induce an accumulation of uncharged To sustain unchecked proliferation, cancer cells evolve to alter
tRNAs that bind to and activate the protein kinase general control translational output by both increasing overall protein synthesis and
non-depressible 2 (GCN2). GCN2 is an essential kinase in the integrated modulating specific mRNA programs38,109. Consistent with this, an
stress response, a short-lived pathway that is necessary for maintain- unbiased analysis of The Cancer Genome Atlas data found that different
ing homeostasis or enabling apoptosis during prolonged or excessive cancers exhibited distinct patterns of aaRS expression110. In a recent
exposure to cellular stress. GCN2 phosphorylates eukaryotic initiation study, LARS1 levels were found to be reduced in breast cancer tumours
factor 2 (eIF2) on its α-subunit at Ser51, causing global inhibition of and cell lines compared with normal tissues and untransformed mam-
protein synthesis and selective upregulation of the translation of acti- mary epithelial cells103. Moreover, LARS1 levels became repressed
vating transcription factor 4 (ATF4), which in turn induces expression during the oncogene-mediated transformation of normal mammary
of adaptive stress response genes36,100. Another, largely underappreci- epithelial cells. Using a genetic mouse model of breast cancer, mono-
ated consequence of amino acid starvation in cancer was described allelic deletion of Lars1 in mouse mammary glands enhanced tumour
to be tRNA misacylation (Fig. 3). The cytokine IFNγ has been found to formation and proliferation. Mechanistically, LARS1 depletion caused
trigger the depletion of the essential amino acid tryptophan (W), which downregulation of select leucine tRNAs. Lower tRNALeu levels inhibited
in turn caused tryptophanyl-tRNA synthetase 1 (WARS1) to erroneously protein synthesis in a codon-dependent manner, which resulted in
mischarge tRNATrp with phenylalanine (F)101. The authors referred to the repression of multiple growth-suppressive genes, including epithelial
W > F containing peptides as ‘substitutants’, as they are not genetically membrane protein 3 (EMP3) and γ-glutamyltransferase 5 (GGT5)103.
encoded, and used functional reporter assays and large-scale proteom- Intriguingly, this study has also revealed that specific isoacceptors
ics to reveal their existence in multiple cancer types. Interestingly, the may act as suppressors of oncogenic transformation, as certain leucine
W > F substitutants were enriched in tumours compared with matched tRNAs were observed to become downregulated in breast cancer cells
adjacent normal tissues, and they could diversify the antigens pre- compared with untransformed human breast epithelial MCF10A
sented by HLA class I at the cell surface of cancer cells to activate T cell cells. Moreover, depletion of tRNALeuCAG has been found to enhance the
responses101. This study highlights the effects of short-term amino acid soft agar colony formation of oncogene-transformed MCF10A cells
starvation on the cancer proteome. relative to controls103 (Fig. 2). Another recent study has also described a
In a recent study, long-term effects of specific amino acid pertur- canonical role for an aaRS in cancer. The authors found that valyl-tRNA
bations in cancer were analysed. Computational analysis of tumour synthetase 1 (VARS1) became upregulated by NOTCH1 signalling in
genome sequencing data has revealed that human cancers are enriched T cell acute lymphoblastic leukaemia. Depletion of VARS1 impaired
in codon-switching events that result in a loss of arginine codons, the translation of mRNAs encoding mitochondrial electron transport
especially in colorectal and stomach cancers102. Limiting arginine chain complex I subunits, resulting in reduced complex I formation,
availability in vitro to physiologically low levels found in such human impaired mitochondrial activity and reduced cellular fitness111. Over-
tumours caused several colorectal cancer cell lines to mutate arginine all, these studies have suggested that aaRSs can become modulated
codons to other codons, thereby ‘losing’ arginine codons. A simi- in a cancer context-dependent manner and selectively modulate
lar arginine codon-switching evolution has also been demonstrated codon-dependent translational gene networks during tumorigenesis.
in vivo as patient-derived colorectal cancer xenografts that underwent
several rounds of liver metastasis selection had significantly more Functions of tRNA fragments in cancer
arginine codon-switching mutations than the parental tumours. The An active area of research that has flourished in recent years relates to
authors note that the liver is a common site of metastasis for colorec- the non-canonical roles of tRNA-derived fragments (tRFs; also referred
tal and gastric cancers, and that it has a high expression of arginase, to as tRNA-derived small RNAs). tRFs are small, single-stranded RNA
the arginine-degrading enzyme. Mechanistically, arginine restriction molecules derived from mature or precursor forms of tRNAs112 that
both perturbed nucleotide synthesis and caused an acute depletion were first detected in the urine of patients with cancer in the 1970s113,114.
of tRNAArg. This caused ribosome stalling at select arginine codons Initially considered inconsequential products of degradation, tRFs
and a proteomic shift from arginine-rich to arginine-low proteins102. are now established as bona fide biologically active small non-coding
Interestingly, the acute degradation of tRNAArg seems to be partially due RNAs implicated in various biological processes112. High-throughput
to loss of aminoacylation (Fig. 3), as depleting arginyl-tRNA synthetase 1 sequencing of small RNAs has afforded the identification of tRFs
(RARS1) reduced tRNAArg levels but no other tRNA species102, in line across all domains of life115 and the assembly of databases curating
with previous work showing that depletion of leucyl-tRNA synthetase 1 tRFs identified in next-generation sequencing cancer data116,117.
(LARS1) reduced the charging of certain leucyl-tRNAs and their overall Depending on their biogenesis pathways, tRFs can be broadly
mature tRNA levels103. The pathway responsible for the degradation classified into six distinct classes: 5′ or 3′ tRNA halves (tiRNA), tRF-1,
of the uncharged tRNAs requires further study, as do the downstream tRF-5, tRF-3 and internal tRFs (i-tRFs)112. The tRNA halves were the first
pathways engaged in this response. tRFs characterized in a model organism118 and are generally associated
with responses to various cellular stresses such as oxidative stress and
Roles of aminoacyl-tRNA synthetases in cancer amino acid starvation119. The tRNA halves are generated by cleavage
aaRSs are evolutionarily ancient enzymes that charge tRNAs with within the anticodon loop by angiogenin99,120, RNase L121 or RNase 1
their cognate amino acids and are thus central to the canonical (ref. 122). tRF-1s are generated during tRNA maturation from the
function of tRNAs in protein synthesis. Owing to their essential 3′ trailer sequences of pre-tRNAs following cleavage by RNase Z123.
roles in translation, mutations in aaRSs are associated with various The mechanism underlying the generation of the remaining classes of

Nature Reviews Cancer


Review article

tRFs is still largely unknown112. The functions and modes of action Tumour-suppressive tRFs
of tRFs are very diverse and have been reviewed elsewhere112,124–126. Several tRFs have been identified in various cancers as tumour sup-
Advancements in small-RNA sequencing and computational biol- pressive, often interacting with Argonaute proteins to engage the RNAi
ogy tools have enabled efforts aimed at identifying specific tRFs or machinery and inhibit the translation of target mRNAs (Fig. 4a). CU1276,
tRF signatures that could be used as biomarkers for various cancers a tRF-3 derived from tRNAGlyGCC, has been observed to be downregulated
(Table 2). Here we summarize studies that have used functional assays in B cell lymphoma cells and DLBCL and found to target the replication
to characterize tRF effects on cancer progression and/or their molecu- protein A1 (RPA1) mRNA that encodes an essential DNA replication and
lar mechanisms of action. On a technical note, tRFs are lowly abundant DNA repair protein128. Ts-3676 has been reported to be downregulated
compared with their parental tRNAs127 and thus specifically deplet- in samples from patients with chronic lymphocytic leukaemia and has
ing tRFs generally minimally impacts the levels of the parental tRNA been found to target the 3′ untranslated region of the TCL1A oncogene
molecules that they arise from. Nevertheless, care should be taken mRNA, reducing its translation. Interestingly, even though ts-3676
in interpreting studies that do not control for the mature tRNA levels interacted with AGO1 and AGO2, it also interacted with Piwi-like pro-
and/or lack additional mechanistic insights into the tRF molecular tein 2 (PIWIL2), suggesting that it might control gene expression
function. through direct site-specific DNA methylation via the PIWI pathway129,130.

Table 2 | Studies on tRNA-derived fragments as cancer biomarkers

Cancer type Identity of tRFs Sample sizes Sample type Ref.

Bladder cancer tRF-5: 5′-tRF Lys


CTT Discovery: 230 bladder cancers Tissue biopsies 154
Validation: 412 TCGA bladder cancers
Breast cancer tRF-1: tRF-Arg-CCT-017, tRF-Gly-CCC-001 Discovery: 8 breast cancers versus 4 controls Plasma samples 155
tRNA half: tiRNAPheGAA003 Validation: 120 breast cancers versus 112 controls
Thirty differentially expressed tRFs (5′ halves; 16 paired breast cancers and normal tissues (6 for initial Tissue biopsies 156
tRF-5, tRF-3, i-tRF) discovery, 10 for validation)
Trastuzumab-resistant i-tRFs from tRNACysGCA: tRF-30-JZOYJE22RR33 57 HER2+ breast cancers previously treated with Serum samples 157
HER2+ breast cancer and tRF-27-ZDXPHO53KSN trastuzumab
Chronic lymphocytic tRF-1, i-tRF, 5′-tRF 20 patients with chronic lymphocytic leukaemia Leukaemia or blood 158
leukaemia and 6 healthy donors samples
Clear cell renal cell 5′-tiRNA4ValAAC Discovery: 18 paired clear cell renal cell carcinomas Tissue biopsies 159
carcinoma and normal tissues
Validation: 118 clear cell renal cell carcinomas and
74 normal tissues
Colorectal cancer 5′-tRFGlyGCC (tRF-35-PNR8YP9LON4VN1) Discovery: 3 colorectal cancers versus 3 healthy controls Plasma samples 160
Validation: 105 colorectal cancers and 90 healthy controls
tRF-3s: tRF-22-WB86Q3P92 (from tRNAThrAGT), Discovery: 104 paired colorectal cancers and normal Tissue biopsies 161
tRF-22-WE8SPOX52 (from tRNAGlyGCC), tissues
tRF-22-WE8S68L52 (from tRNAGlyCCC), Validation: TCGA data, 603 primary tumours and
tRF-18-8R1546D2 (from tRNAAlaAGC) 11 adjacent normal tissues
Non-small-cell lung tRNA halves, potentially other Discovery: 30 pairs of non-small cell lung cancers and Tissue biopsies, 162
cancer normal tissues serum samples
Validation: 60 pairs of non-small cell lung cancers and
normal tissues; 167 serum samples (34 normal and 133
non-small-cell lung cancers)
tRF-5: tRF-31-79MP9P9NH57SD 96 non-small-cell lung cancer, 96 healthy controls and Serum samples 163
(from tRNAValAAC/CAC) 20 pairs of non-small-cell lung cancer serum samples
pre-surgery and post-surgery
Gastric cancer i-tRF: tRF-31-U5YKFN8DYDZDD Discovery: 3 paired gastric cancer samples Tissue biopsies, 164
(from tRNAValTAC) Validation: 24 pairs of gastric cancer samples; sera from serum samples
111 patients with gastric cancer, 48 gastritis and 89 healthy
controls; 28 pre-operative and post-operative patients with
gastric cancer
Ovarian cancer i-tRF-Gly-GCC Screening: 98 ovarian cancers Tissue biopsies 165
Validation: 100 ovarian cancers
Prostate cancer All tRF classes; ratio of tRF-315 (from Discovery: 31 prostate cancer and 8 non-malignant Tissue biopsies 166
tRNALysCTT) to tRF-544 (from tRNAPheGAA) samples
proposed as a biomarker Validation: 65 prostate cancers (cohort 1) and 104 prostate
cancers (cohort 2)
i-tRF, internal transfer RNA-derived fragment; TGCA, The Cancer Genome Atlas; tRNA, transfer RNA.

Nature Reviews Cancer


Review article

a 3′ b c mTOGs tRF-21
3′ YBX1 YBX1
5′
tRF
tRF-Glu-YUC tRF-Asp-GUC
PAIP1 DDX17
mRNA mRNA YBX1 YBX1 PABPC1 hnRNPL
AGO target degradation S52
tRF-Gly-UCC tRF-Tyr-GUA

Tumour-suppressive tRFs
Destabilization of growth- PAIP1 hnRNPL
tRFs mRNAs PABPC1 DDX17
promoting mRNAs

CU1276 RPA1
ts-3676 TCL1A
tRF-24-V29K9UV3IU GPR78 Translation of Altered splicing
5′ PES mRNAs CASP9, MACROH2A1
Oncogenic or tumorigenic tRFs
tRF-3017A NELL2
tRF-3019a FBX047 Breast cancer metastasis MDS progression to AML Pancreatic cancer progression

d e NCL NCL
RBM17
5′-tiRNA Gly
GCC
UHM 5′-tRF
Cys
5′-tRFCys
5′ Oligomerization Breast
RB

cancer
5′
M

RBM17 metastasis
17

Altered
splicing Thyroid 5′ mRNAs 3′ 3′
3′ 3′
cancer 3′ 3′
Target mRNAs
Spliceosome progression
PAFAH1B1 and MTHFD1L mRNA stabilization

Fig. 4 | Examples of oncogenic and tumour-suppressive transfer RNA-derived (PES) in their 5′ untranslated regions and thus could inhibit progression of
fragments in cancer. a, Transfer RNA (tRNA)-derived fragments (tRFs) can myelodysplastic syndrome (MDS) to acute myeloid leukaemia (AML) (left).
engage the cellular RNA interference machinery and inhibit the translation In pancreatic cancer (right), tRF-21-VBY9PYKHD binds to Ser52 on heterogeneous
of target messenger RNA (mRNA) by inducing mRNA degradation. Potential nuclear ribonucleoprotein L (hnRNPL) and prevents its phosphorylation at
tumour-suppressive and tumorigenic microRNA-like tRFs and their target genes that residue, which inhibits the formation of an RNA splicing complex with
are shown. b, tRFs can bind to RNA-binding proteins (RBPs), such as the Y-box DEAD-box helicase 17 (DDX17). This results in altered splicing of select
binding protein 1 (YBX1), and segregate them away from their target mRNAs, mRNAs (part c). Examples of tumorigenic tRFs modulating protein–protein
which, for example, results in the destabilization of growth-promoting target interactions (parts d and e). The 5′-tiRNAGlyGCC may promote thyroid cancer
mRNAs and suppression of breast cancer metastasis. c–e, tRFs can modulate the progression through altered splicing (part d) and 5′-tRFCys may promote
interactions of RBPs with other proteins. Examples of tumour-suppressive tRFs breast cancer metastasis through mRNA stabilization (part e). AGO,
modulating protein–protein interactions include pseudouridylated mini tRNAs Argonaute; CASP9, caspase 9; MACROH2A1, macroH2A.1 histone; MTHFD1L,
containing a 5′ terminal oligoguanine (mTOGs, a type of tRF-5s) that inhibit the methylenetetrahydrofolate dehydrogenase (NADP+ dependent) 1-like; NCL,
binding of poly(A)-binding protein 1 (PABPC1) to the translational co-activator nucleolin; PAFAH1B1, platelet-activating factor acetylhydrolase 1b regulatory
poly(A)-binding protein interacting protein 1 (PAIP1), which represses subunit 1; RBM17, RNA-binding motif protein 17; UHM, U2AF homology motif.
the translation of mRNAs containing pyrimidine (Y)-enriched sequences

Finally, the knockdown of tRF-24-V29K9UV3IU, a tRF-5 downregulated (Fig. 3). In highly metastatic breast cancer cells, the hypoxic induction
in gastric cancer tumours, promoted the growth and metastasis of of a subset of tRFs was blunted, and the group of four tRFs derived
tumour xenografts of MKN-45 gastric cancer cells, whereas overexpres- from tRNAGluYTC, tRNAAspGTC, tRNAGlyTCC and tRNATyrGTA was repressed in
sion of the tRF inhibited cell proliferation, migration and invasion131. The highly metastatic cells relative to poorly metastatic cells97. These tRFs
authors found that tRF-24-V29K9UV3IU targeted, in a miRNA-like man- inhibited metastasis when upregulated and, conversely, promoted
ner, the G protein-coupled receptor GPR78, which had previously been metastasis when inhibited in breast cancer cells. Mechanistically,
shown to promote malignant phenotypes in lung cancer132. Epistasis the tRFs derived from tRNAGluYTC, tRNAAspGTC, tRNAGlyTCC and tRNATyrGTA
experiments have shown that altered GPR78 levels accounted partially bound to and competitively inhibited the binding of Y-box binding
for the phenotypes caused by overexpression of tRF-24-V29K9UV3IU131. protein 1 (YBX1) to a set of growth-promoting and oncogenic mRNAs.
Another major mode of action for tRFs in cancer involves interac- Because YBX1 stabilized those transcripts, tRF binding to YBX1 led
tions with RNA-binding proteins (RBPs). RBPs regulate gene expression to their destabilization and suppressed metastasis97. Another study
post-transcriptionally through a diverse set of mechanisms, such as has also demonstrated that fragmentation of tRNAGlu can give rise to
modulating the translation of specific mRNAs133. A tRF can bind to an a tumour-suppressive tRF in breast cancer. The tRF identified in this
RBP and competitively inhibit the binding of the RBP to its mRNA tar- study, termed tRF3E, was shown to bind to nucleolin and was suspected
gets, affecting the stability of the respective mRNAs (Fig. 4b). Interest- of inhibiting its binding to its downstream target mRNAs134.
ingly, hypoxia has been found to induce specific tRNA fragmentation in tRFs can also inhibit the interaction of RBPs with other proteins
breast cancer cells that was functionally linked to metastatic potential97 (Fig. 4c). Recently, a class of pseudouridylated tRF-5s called mini tRNAs

Nature Reviews Cancer


Review article

containing a 5′ terminal oligoguanine (mTOGs) have been shown to (Pafah1b1) and methylenetetrahydrofolate dehydrogenase (NADP+
regulate translation initiation, implicating them in myelodysplastic dependent) 1-like (Mthfd1l). Further experiments confirmed that
syndrome progression135. Pseudouridylated mTOGs have been found both metabolic enzymes act downstream of 5′-tRFCysGCA to mediate its
to bind to poly(A)-binding protein 1 (PABPC1) and prevent its interac- pro-survival and pro-metastatic activities. Overall, this study revealed
tion with translational co-activator poly(A)-binding protein interacting that a tRF can act in a gain-of-function manner to promote the forma-
protein 1 (PAIP1), which resulted in repression of mRNA transcripts con- tion of a stabilizing ribonucleoprotein complex on mRNAs, resulting
taining pyrimidine-enriched sequences in their 5′ untranslated regions. in pro-tumorigenic phenotypes141.
This translational repression prevented aberrant protein synthesis that
could drive myelodysplastic syndrome progression to acute myeloid leu- Conclusions and future perspectives
kaemia, although the identity of the repressed oncogenic translational The past 15 years have brought tRNAs into focus as dynamic and critical
programs has not yet been determined135,136. Finally, in pancreatic cancer, regulators of cancer progression. Even though imperfect (Box 1), tech-
an i-tRF derived from tRNAGlyGCC, tRF-21-VBY9PYKHD, has been found nological advances have uncovered global tRNA alterations in tumours
to be downregulated in tumours compared with normal tissues and in that suggest cancer cells optimize their tRNA pools to match the codon
later-stage disease versus earlier stages137. Extensive gain-of-function usage of their transcriptomes, allowing for efficient translation. Fur-
and loss-of-function experiments in vivo and in vitro have supported the thermore, studies have shown that individual tRNAs can have causal
conclusion that tRF-21-VBY9PYKHD is a tumour suppressor in pancreatic roles in driving cancer progression via modulation of distinct gene
cancer. The authors showed that tRF-21-VBY9PYKHD interacted with expression networks in a codon-dependent manner. These findings
heterogeneous nuclear ribonucleoprotein L (hnRNPL) at residue Ser52, open up new questions such as whether the tRNA abundance changes
which blocked phosphorylation of hnRNPL, inhibiting its formation of an during cancer progression disproportionally affect the translation effi-
RNA splicing complex with DEAD-box helicase 17 (DDX17). This resulted ciency of genes enriched in rare codons, especially given the evidence
in altered splicing of mRNAs encoding caspase 9 and macroH2A.1 his- that the enrichment of rare codons in the common oncogene KRAS
tone, inhibiting the formation of the anti-apoptotic caspase 9b and the influence its expression142, oncogenesis143,144 and potentially treatment
pro-invasive mH2A.1.2 isoforms. The authors also showed that tRF-21 for- responses145. Moreover, the reports so far have only been restricted
mation is mediated by serine and arginine-rich splicing factor 5 (SRSF5), to a handful of cancer types, indicating that the systematic analyses
which is downregulated following treatment with the inflammatory of tRNA levels using dedicated tRNA profiling techniques in different
cytokines LIF IL-6 family cytokine (LIF) or IL-6 (ref. 137). These studies cancer types are needed (Box 1). Gain-of-function and loss-of-function
have revealed that tRF binding to RBPs and altering their interactions is experiments are further needed to then pinpoint the tRNA changes that
a major mechanism of action for this class of non-coding RNAs. functionally impact cancer progression in different cancer types (Box 1).
The tRNA-related basic research discoveries open new avenues
Tumour-promoting tRFs for diagnostic and therapeutic interventions for patients with can-
Several tRFs can downregulate the expression of potential tumour sup- cer. For example, there is increasing interest in using small-molecule
pressors. In gastric cancer, tRF-3017A and tRF-3019a have been found to inhibitors to regulate the activity of aaRSs in human disease, including
be upregulated in tumours138,139. Experimentally, modulating the levels cancer146–148. Moreover, tRNAs themselves could be delivered as thera-
of each of the tRFs impacted cell migration and invasion in vitro138,139. pies, as illustrated by recent proof-of-concept studies. For example,
The authors then showed that tRF-3017A targeted neural EGFL-like 2 transgenic overexpression of tRNAGly rescued pathogenic phenotypes
(NELL2)138, whereas tRF-3019a targeted F-box protein 47 (FBXO47)139, in Drosophila and murine models of Charcot–Marie–Tooth peripheral
and suggested that repression of these potential tumour suppressors neuropathy149. Other studies have illustrated different methods to
mediated the phenotypes of these tRFs, respectively (Fig. 4a). deliver tRNAs in mice, focusing on suppressor tRNAs150,151. A suppres-
Oncogenic tRFs can also mediate their effects by regulating sor tRNA carries a mutation in its anticodon that allows it to bind to a
RBPs (Fig. 4d,e). For example, the 5′ tRNA half 5′-tiRNAGlyGCC has been stop codon and incorporate an amino acid in its place, which enables
found to be overexpressed in papillary thyroid carcinoma (PTC)140. readthrough of a premature stop codon introduced by a nonsense
Exogenous expression of 5′-tiRNAGlyGCC increased cell proliferation mutation152. Using an adeno-associated virus system to deliver a sup-
and cell migration in vitro in PTC cells, whereas its depletion had pressor tRNATyr, readthrough of the mutant Idua gene (Idua(W401X);
the opposite effects. Consistent with the in vitro data, knockdown TGG→TAG) was successfully induced and the pathogenic phenotypes
of 5′-tiRNAGlyGCC inhibited tumour growth in xenograft assays using were partially rescued in a mouse model that recapitulates the human
a PTC cell line140. Mechanistically, 5′-tiRNAGlyGCC bound the splicing lysosomal storage disease mucopolysaccharidosis type I150. Another
protein RNA-binding motif protein 17 (RBM17) via its U2AF homo­ proof-of-concept paper used lipid nanoparticles to successfully deliver
logy motif domain, resulting in RBM17 stabilization and its translo- functional suppressor tRNAs in mice151. An optimistic therapeutic
cation to the nucleus. Overexpression of 5′-tiRNAGlyGCC in a PTC cell outlook inspired by these papers is the delivery of tRNAs to restore the
line changed the splicing patterns of genes associated with onco- levels of tumour-suppressive tRNAs in human cancer.
genic signatures, indicating that 5′-tiRNAGlyGCC exerts its function in The characterization of tRFs as bona fide functional non-coding
part by RBM17-mediated alternative splicing140. Another tRF that small RNAs coupled with sequencing of patient samples led to the
exerts its function via RBPs is the 5′-tRFCysGCA, which is upregulated identification of many tRFs with functional roles in cancer. The isola-
during breast cancer progression and associated with reduced sur- tion and sequencing of small RNAs from liquid biopsies can aid the
vival outcomes in patients with breast cancer141. Depleting 5′-tRFCysGCA discovery of tRF signatures that could serve as biomarkers indicative
inhibited breast cancer cell survival and metastasis in xenograft of different cancer types and even cancer stages. Moreover, tRFs that
models. Mechanistically, 5′-tRFCysGCA bound to nucleolin, leading manifest their effects on cancer progression by regulating the activ-
to its oligomerization and stabilization of mRNAs bound to nucleolin, ity of RBPs could be potential therapeutic targets using technologies
such as platelet activating factor acetylhydrolase 1b regulatory subunit 1 such as locked-nucleic acids153. Locked-nucleic acids could also be used

Nature Reviews Cancer


Review article

to target mature tRNAs that promote tumour growth or metastasis. 29. Chan, P. P., Lin, B. Y., Mak, A. J. & Lowe, T. M. tRNAscan-SE 2.0: improved detection and
functional classification of transfer RNA genes. Nucleic Acids Res. 49, 9077–9096 (2021).
These are just a few of the potential translational applications of the This resource paper, together with Chan et al. (2016), describes the GtRNAdb
findings related to tRNA in cancer. We envision that even more appli- database of predicted tRNA sequences.
cations and potential for therapeutic intervention will be discovered 30. Pavon-Eternod, M. et al. tRNA over-expression in breast cancer and functional
consequences. Nucleic Acids Res. 37, 7268–7280 (2009).
as we continue to uncover how tRNA biology is functionally linked to This paper is one of the first to attempt global profiling of tRNA levels in cancer using
cancer progression. microarrays.
31. Pavon-Eternod, M., Gomes, S., Rosner, M. R. & Pan, T. Overexpression of initiator
methionine tRNA leads to global reprogramming of tRNA expression and increased
Published online: xx xx xxxx proliferation in human epithelial cells. RNA 19, 461–466 (2013).
32. Earnest-Noble, L. B. et al. Two isoleucyl tRNAs that decode synonymous codons
References divergently regulate breast cancer metastatic growth by controlling translation
1. Hoagland, M. B., Zamecnik, P. C. & Stephenson, M. L. Intermediate reactions in protein of proliferation-regulating genes. Nat. Cancer 3, 1484–1497 (2022).
biosynthesis. Biochim. Biophys. Acta 24, 215–216 (1957). 33. Taylor, M. W., Granger, G. A., Buck, C. A. & Holland, J. J. Similarities and differences
2. Ogata, K. & Nohara, H. The possible role of the ribonucleic acid (RNA) of the pH 5 enzyme among specific tRNA’s in mammalian tissues. Proc. Natl Acad. Sci. USA 57, 1712–1719
in amino acid activation. Biochim. Biophys. Acta 25, 659–660 (1957). (1967).
3. Holley, R. W. An alanine-dependent, ribonuclease-inhibited conversion of AMP to 34. Gallo, R. C. Transfer RNA’s in human leukemia. J. Cell Physiol. 74, 149–153 (1969).
ATP, and its possible relationship to protein synthesis. J. Am. Chem. Soc. 79, 658–662 35. Baliga, B. S., Borek, E., Weinstein, I. B. & Srinivasan, P. R. Differences in the transfer
(1957). RNA’s of normal liver and Novikoff hepatoma. Proc. Natl Acad. Sci. USA 62, 899–905
This study, together with Hoagland et al. (1957) and Ogata et al. (1957), is one of the (1969).
original studies that discovered tRNAs. 36. Sonenberg, N. & Hinnebusch, A. G. Regulation of translation initiation in eukaryotes:
4. Söll, D. & RajBhandary, U. L. tRNA: Structure, Biosynthesis, and Function (Wiley, 1995). mechanisms and biological targets. Cell 136, 731–745 (2009).
5. Suzuki, T. The expanding world of tRNA modifications and their disease relevance. 37. Birch, J. et al. The initiator methionine tRNA drives cell migration and invasion leading
Nat. Rev. Mol. Cell Biol. 22, 375–392 (2021). to increased metastatic potential in melanoma. Biol. Open 5, 1371–1379 (2016).
6. Suzuki, T., Nagao, A. & Suzuki, T. Human mitochondrial tRNAs: biogenesis, function, 38. Ruggero, D. Translational control in cancer etiology. Cold Spring Harb. Perspect. Biol. 5,
structural aspects, and diseases. Annu. Rev. Genet. 45, 299–329 (2011). a012336 (2013).
This is a comprehensive review on human mt-tRNAs. 39. Silvera, D., Formenti, S. C. & Schneider, R. J. Translational control in cancer. Nat. Rev.
7. Orellana, E. A., Siegal, E. & Gregory, R. I. tRNA dysregulation and disease. Nat. Rev. Genet. Cancer 10, 254–266 (2010).
23, 651–664 (2022). 40. Clarke, C. J. et al. The initiator methionine tRNA drives secretion of type II collagen from
8. Holley, R. W. et al. Structure of a ribonucleic acid. Science 147, 1462–1465 (1965). stromal fibroblasts to promote tumor growth and angiogenesis. Curr. Biol. 26, 755–765
This paper describes the first sequence of a tRNA. (2016).
9. Kim, S. H. et al. Three-dimensional tertiary structure of yeast phenylalanine transfer RNA. 41. Yang, J. et al. SOX4-mediated repression of specific tRNAs inhibits proliferation of human
Science 185, 435–440 (1974). glioblastoma cells. Proc. Natl Acad. Sci. USA 117, 5782–5790 (2020).
10. Cramer, F., Erdmann, V. A., von der Haar, F. & Schlimme, E. Structure and reactivity 42. Wang, B. et al. miR-34a directly targets tRNAiMet precursors and affects cellular
of tRNA. J. Cell Physiol. 74, 163–178 (1969). proliferation, cell cycle, and apoptosis. Proc. Natl Acad. Sci. USA 115, 7392–7397
11. Ramsay, E. P. & Vannini, A. Structural rearrangements of the RNA polymerase III (2018).
machinery during tRNA transcription initiation. Biochim. Biophys. Acta Gene Regul. 43. Shigematsu, M. et al. YAMAT-seq: an efficient method for high-throughput sequencing
Mech. 1861, 285–294 (2018). of mature transfer RNAs. Nucleic Acids Res. 45, e70 (2017).
12. Phizicky, E. M. & Hopper, A. K. tRNA biology charges to the front. Genes Dev. 24, 44. Behrens, A., Rodschinka, G. & Nedialkova, D. D. High-resolution quantitative profiling of
1832–1860 (2010). tRNA abundance and modification status in eukaryotes by mim-tRNAseq. Mol. Cell 81,
13. Rak, R. et al. Dynamic changes in tRNA modifications and abundance during 1802–1815.e7 (2021).
T cell activation. Proc. Natl Acad. Sci. USA https://doi.org/10.1073/pnas.2106556118 45. Gogakos, T. et al. Characterizing expression and processing of precursor and mature
(2021). human tRNAs by hydro-tRNAseq and PAR-CLIP. Cell Rep. 20, 1463–1475 (2017).
14. Torrent, M., Chalancon, G., de Groot, N. S., Wuster, A. & Madan Babu, M. Cells alter their 46. Lucas, M. C. et al. Quantitative analysis of tRNA abundance and modifications by
tRNA abundance to selectively regulate protein synthesis during stress conditions. nanopore RNA sequencing. Nat. Biotechnol. https://doi.org/10.1038/s41587-023-01743-6
Sci. Signal. 11, eaat6409 (2018). (2023).
15. Dittmar, K. A., Goodenbour, J. M. & Pan, T. Tissue-specific differences in human transfer 47. Iben, J. R. & Maraia, R. J. tRNA gene copy number variation in humans. Gene 536,
RNA expression. PLoS Genet. 2, e221 (2006). 376–384 (2014).
16. Pinkard, O., McFarland, S., Sweet, T. & Coller, J. Quantitative tRNA-sequencing uncovers 48. Darrow, E. M. & Chadwick, B. P. A novel tRNA variable number tandem repeat at human
metazoan tissue-specific tRNA regulation. Nat. Commun. 11, 4104 (2020). chromosome 1q23.3 is implicated as a boundary element based on conservation of a
17. Kimura, S., Dedon, P. C. & Waldor, M. K. Comparative tRNA sequencing and RNA mass CTCF motif in mouse. Nucleic Acids Res. 42, 6421–6435 (2014).
spectrometry for surveying tRNA modifications. Nat. Chem. Biol. 16, 964–972 (2020). 49. Parisien, M., Wang, X. & Pan, T. Diversity of human tRNA genes from the 1000-Genomes
18. Gingold, H. et al. A dual program for translation regulation in cellular proliferation Project. RNA Biol. 10, 1853–1867 (2014).
and differentiation. Cell 158, 1281–1292 (2014). 50. Berg, M. D. et al. Targeted sequencing reveals expanded genetic diversity of human
This paper classifies ‘proliferation-specific’ and ‘differentiation-specific’ tRNAs. transfer RNAs. RNA Biol. 16, 1574–1585 (2019).
19. Oler, A. J. et al. Human RNA polymerase III transcriptomes and relationships to Pol II 51. Thornlow, B. P. et al. Transfer RNA genes experience exceptionally elevated mutation
promoter chromatin and enhancer-binding factors. Nat. Struct. Mol. Biol. 17, 620–628 rates. Proc. Natl Acad. Sci. USA 115, 8996–9001 (2018).
(2010). 52. Ishimura, R. et al. Ribosome stalling induced by mutation of a CNS-specific tRNA causes
20. Barski, A. et al. Pol II and its associated epigenetic marks are present at Pol III-transcribed neurodegeneration. Science 345, 455–459 (2014).
noncoding RNA genes. Nat. Struct. Mol. Biol. 17, 629–634 (2010). This is one of the first reports of a tRNA mutation causing a tissue-specific phenotype
21. Goodarzi, H. et al. Modulated expression of specific tRNAs drives gene expression and in a mouse model.
cancer progression. Cell 165, 1416–1427 (2016). 53. Schoenmakers, E. et al. Mutation in human selenocysteine transfer RNA selectively
22. Boccaletto, P. et al. MODOMICS: a database of RNA modification pathways. 2021 update. disrupts selenoprotein synthesis. J. Clin. Invest. 126, 992–996 (2016).
Nucleic Acids Res. 50, D231–D235 (2022). 54. Santos, M. et al. Codon misreading tRNAs promote tumor growth in mice. RNA Biol. 15,
23. Zhang, W., Foo, M., Eren, A. M. & Pan, T. tRNA modification dynamics from individual 773–786 (2018).
organisms to metaepitranscriptomics of microbiomes. Mol. Cell 82, 891–906 (2022). 55. Doran, J. L., Bingle, W. H. & Roy, K. L. Two human genes encoding tRNA. Gene 65,
This article includes a systematic analysis of current tRNA profiling methods. 329–336 (1988).
24. Chujo, T. & Tomizawa, K. Human transfer RNA modopathies: diseases caused by 56. Sprague, K. U., Larson, D. & Morton, D. 5′ Flanking sequence signals are required for
aberrations in transfer RNA modifications. FEBS J. 288, 7096–7122 (2021). activity of silkworm alanine tRNA genes in homologous in vitro transcription systems.
25. Barbieri, I. & Kouzarides, T. Role of RNA modifications in cancer. Nat. Rev. Cancer 20, Cell 22, 171–178 (1980).
303–322 (2020). 57. Park, J. L. et al. Epigenetic regulation of noncoding RNA transcription by mammalian RNA
26. Clark, W. C., Evans, M. E., Dominissini, D., Zheng, G. & Pan, T. tRNA base methylation polymerase III. Epigenomics 9, 171–187 (2017).
identification and quantification via high-throughput sequencing. RNA 22, 1771–1784 58. White, R. J. Transcription by RNA polymerase III: more complex than we thought.
(2016). Nat. Rev. Genet. 12, 459–463 (2011).
27. Crick, F. H. Codon–anticodon pairing: the wobble hypothesis. J. Mol. Biol. 19, 548–555 59. Larminie, C. G. et al. Mechanistic analysis of RNA polymerase III regulation by the
(1966). retinoblastoma protein. EMBO J. 16, 2061–2071 (1997).
28. Chan, P. P. & Lowe, T. M. GtRNAdb 2.0: an expanded database of transfer RNA genes 60. Cairns, C. A. & White, R. J. p53 is a general repressor of RNA polymerase III transcription.
identified in complete and draft genomes. Nucleic Acids Res. 44, D184–D189 (2016). EMBO J. 17, 3112–3123 (1998).

Nature Reviews Cancer


Review article

61. Crighton, D. et al. p53 represses RNA polymerase III transcription by targeting TBP and 95. Schwenzer, H. et al. Oxidative stress triggers selective tRNA retrograde transport in
inhibiting promoter occupancy by TFIIIB. EMBO J. 22, 2810–2820 (2003). human cells during the integrated stress response. Cell Rep. 26, 3416–3428.e5 (2019).
62. Gomez-Roman, N., Grandori, C., Eisenman, R. N. & White, R. J. Direct activation of RNA 96. Thompson, D. M., Lu, C., Green, P. J. & Parker, R. tRNA cleavage is a conserved response
polymerase III transcription by c-Myc. Nature 421, 290–294 (2003). to oxidative stress in eukaryotes. RNA 14, 2095–2103 (2008).
63. Kenneth, N. S. et al. TRRAP and GCN5 are used by c-Myc to activate RNA polymerase III 97. Goodarzi, H. et al. Endogenous tRNA-derived fragments suppress breast cancer
transcription. Proc. Natl Acad. Sci. USA 104, 14917–14922 (2007). progression via YBX1 displacement. Cell 161, 790–802 (2015).
64. Willis, I. M. & Moir, R. D. Signaling to and from the RNA polymerase III transcription and 98. Huh, D. et al. A stress-induced tyrosine-tRNA depletion response mediates codon-based
processing machinery. Annu. Rev. Biochem. 87, 75–100 (2018). translational repression and growth suppression. EMBO J. 40, e106696 (2021).
65. Raha, D. et al. Close association of RNA polymerase II and many transcription factors with 99. Yamasaki, S., Ivanov, P., Hu, G. F. & Anderson, P. Angiogenin cleaves tRNA and promotes
Pol III genes. Proc. Natl Acad. Sci. USA 107, 3639–3644 (2010). stress-induced translational repression. J. Cell Biol. 185, 35–42 (2009).
66. Zhang, J. et al. SOX4 inhibits GBM cell growth and induces G0/G1 cell cycle arrest 100. Masson, G. R. Towards a model of GCN2 activation. Biochem. Soc. Trans. 47, 1481–1488
through Akt–p53 axis. BMC Neurol. 14, 207 (2014). (2019).
67. Smith, D. K., Yang, J., Liu, M. L. & Zhang, C. L. Small molecules modulate chromatin 101. Pataskar, A. et al. Tryptophan depletion results in tryptophan-to-phenylalanine
accessibility to promote NEUROG2-mediated fibroblast-to-neuron reprogramming. Stem substitutants. Nature 603, 721–727 (2022).
Cell Rep. 7, 955–969 (2016). 102. Hsu, D. J. et al. Arginine limitation drives a directed codon-dependent DNA sequence
68. Lee, H., Goodarzi, H., Tavazoie, S. F. & Alarcon, C. R. TMEM2 is a SOX4-regulated gene evolution response in colorectal cancer cells. Sci. Adv. 9, eade9120 (2023).
that mediates metastatic migration and invasion in breast cancer. Cancer Res. 76, 103. Passarelli, M. C. et al. Leucyl-tRNA synthetase is a tumour suppressor in breast
4994–5005 (2016). cancer and regulates codon-dependent translation dynamics. Nat. Cell Biol. 24,
69. Vervoort, S. J., van Boxtel, R. & Coffer, P. J. The role of SRY-related HMG box transcription 307–315 (2022).
factor 4 (SOX4) in tumorigenesis and metastasis: friend or foe? Oncogene 32, 104. Turvey, A. K., Horvath, G. A. & Cavalcanti, A. R. O. Aminoacyl-tRNA synthetases in human
3397–3409 (2013). health and disease. Front. Physiol. 13, 1029218 (2022).
70. Besser, D. et al. DNA methylation inhibits transcription by RNA polymerase III of a tRNA 105. Hyeon, D. Y. et al. Evolution of the multi-tRNA synthetase complex and its role in cancer.
gene, but not of a 5S rRNA gene. FEBS Lett. 269, 358–362 (1990). J. Biol. Chem. 294, 5340–5351 (2019).
71. Bartke, T. et al. Nucleosome-interacting proteins regulated by DNA and histone 106. Wang, J. & Yang, X. L. Novel functions of cytoplasmic aminoacyl-tRNA synthetases
methylation. Cell 143, 470–484 (2010). shaping the hallmarks of cancer. Enzymes 48, 397–423 (2020).
72. Rossello-Tortella, M., Bueno-Costa, A., Martinez-Verbo, L., Villanueva, L. & Esteller, M. 107. Sung, Y., Yoon, I., Han, J. M. & Kim, S. Functional and pathologic association of
DNA methylation-associated dysregulation of transfer RNA expression in human cancer. aminoacyl-tRNA synthetases with cancer. Exp. Mol. Med. 54, 553–566 (2022).
Mol. Cancer 21, 48 (2022). 108. Guo, M. & Schimmel, P. Essential nontranslational functions of tRNA synthetases.
73. Tavtigian, S. V. et al. A candidate prostate cancer susceptibility gene at chromosome 17p. Nat. Chem. Biol. 9, 145–153 (2013).
Nat. Genet. 27, 172–180 (2001). 109. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144,
74. Xu, B., Tong, N., Li, J. M., Zhang, Z. D. & Wu, H. F. ELAC2 polymorphisms and prostate 646–674 (2011).
cancer risk: a meta-analysis based on 18 case-control studies. Prostate Cancer Prostatic 110. Wang, J. et al. Multi-omics database analysis of aminoacyl-tRNA synthetases in cancer.
Dis. 13, 270–277 (2010). Genes 11, 1384 (2020).
75. Lubas, M., Chlebowski, A., Dziembowski, A. & Jensen, T. H. Biochemistry and function of 111. Thandapani, P. et al. Valine tRNA levels and availability regulate complex I assembly in
RNA exosomes. Enzymes 31, 1–30 (2012). leukaemia. Nature 601, 428–433 (2022).
76. Kadaba, S. et al. Nuclear surveillance and degradation of hypomodified initiator tRNAMet 112. Su, Z., Wilson, B., Kumar, P. & Dutta, A. Noncanonical roles of tRNAs: tRNA fragments and
in S. cerevisiae. Genes Dev. 18, 1227–1240 (2004). beyond. Annu. Rev. Genet. 54, 47–69 (2020).
77. Whipple, J. M., Lane, E. A., Chernyakov, I., D’Silva, S. & Phizicky, E. M. The yeast rapid tRNA 113. Borek, E. et al. High turnover rate of transfer RNA in tumor tissue. Cancer Res. 37,
decay pathway primarily monitors the structural integrity of the acceptor and T-stems of 3362–3366 (1977).
mature tRNA. Genes Dev. 25, 1173–1184 (2011). 114. Speer, J., Gehrke, C. W., Kuo, K. C., Waalkes, T. P. & Borek, E. tRNA breakdown products as
78. Alexandrov, A. et al. Rapid tRNA decay can result from lack of nonessential markers for cancer. Cancer 44, 2120–2123 (1979).
modifications. Mol. Cell 21, 87–96 (2006). 115. Kumar, P., Anaya, J., Mudunuri, S. B. & Dutta, A. Meta-analysis of tRNA derived RNA
79. De Zoysa, T. & Phizicky, E. M. Hypomodified tRNA in evolutionarily distant yeasts can fragments reveals that they are evolutionarily conserved and associate with AGO
trigger rapid tRNA decay to activate the general amino acid control response, but with proteins to recognize specific RNA targets. BMC Biol. 12, 78 (2014).
different consequences. PLoS Genet. 16, e1008893 (2020). 116. Pliatsika, V. et al. MINTbase v2.0: a comprehensive database for tRNA-derived fragments
80. Wilusz, J. E., Whipple, J. M., Phizicky, E. M. & Sharp, P. A. tRNAs marked with CCACCA are that includes nuclear and mitochondrial fragments from all the cancer genome atlas
targeted for degradation. Science 334, 817–821 (2011). projects. Nucleic Acids Res. 46, D152–D159 (2018).
81. Orellana, E. A. et al. METTL1-mediated m7G modification of Arg-TCT tRNA drives 117. Yao, D. et al. OncotRF: an online resource for exploration of tRNA-derived fragments
oncogenic transformation. Mol. Cell 81, 3323–3338.e14 (2021). in human cancers. RNA Biol. 17, 1081–1091 (2020).
82. Han, H. et al. N7-methylguanosine tRNA modification promotes esophageal squamous 118. Lee, S. R. & Collins, K. Starvation-induced cleavage of the tRNA anticodon loop in
cell carcinoma tumorigenesis via the RPTOR/ULK1/autophagy axis. Nat. Commun. 13, Tetrahymena thermophila. J. Biol. Chem. 280, 42744–42749 (2005).
1478 (2022). 119. Tao, E. W., Cheng, W. Y., Li, W. L., Yu, J. & Gao, Q. Y. tiRNAs: a novel class of small
83. Chen, J. et al. Aberrant translation regulated by METTL1/WDR4-mediated tRNA noncoding RNAs that helps cells respond to stressors and plays roles in cancer
N7-methylguanosine modification drives head and neck squamous cell carcinoma progression. J. Cell Physiol. 235, 683–690 (2020).
progression. Cancer Commun. 42, 223–244 (2022). 120. Fu, H. et al. Stress induces tRNA cleavage by angiogenin in mammalian cells. FEBS Lett.
84. Dai, Z. et al. N7-methylguanosine tRNA modification enhances oncogenic mRNA 583, 437–442 (2009).
translation and promotes intrahepatic cholangiocarcinoma progression. Mol. Cell 81, 121. Donovan, J., Rath, S., Kolet-Mandrikov, D. & Korennykh, A. Rapid RNase L-driven arrest of
3339–3355.e8 (2021). protein synthesis in the dsRNA response without degradation of translation machinery.
85. Chen, Z. et al. METTL1 promotes hepatocarcinogenesis via m7G tRNA RNA 23, 1660–1671 (2017).
modification-dependent translation control. Clin. Transl. Med. 11, e661 (2021). 122. Nechooshtan, G., Yunusov, D., Chang, K. & Gingeras, T. R. Processing by RNase 1 forms
86. Pandolfini, L. et al. METTL1 promotes let-7 microRNA processing via m7G methylation. tRNA halves and distinct Y RNA fragments in the extracellular environment. Nucleic
Mol. Cell 74, 1278–1290.e9 (2019). Acids Res. 48, 8035–8049 (2020).
87. Cheng, W., Gao, A., Lin, H. & Zhang, W. Novel roles of METTL1/WDR4 in tumor via m7G 123. Lee, Y. S., Shibata, Y., Malhotra, A. & Dutta, A. A novel class of small RNAs: tRNA-derived
methylation. Mol. Ther. Oncolytics 26, 27–34 (2022). RNA fragments (tRFs). Genes Dev. 23, 2639–2649 (2009).
88. Girstmair, H. et al. Depletion of cognate charged transfer RNA causes translational 124. Oberbauer, V. & Schaefer, M. R. tRNA-derived small RNAs: biogenesis, modification,
frameshifting within the expanded CAG stretch in huntingtin. Cell Rep. 3, 148–159 (2013). function and potential impact on human disease development. Genes 9, 607 (2018).
89. Guzy, R. D. et al. Mitochondrial complex III is required for hypoxia-induced ROS 125. Guzzi, N. & Bellodi, C. Novel insights into the emerging roles of tRNA-derived fragments
production and cellular oxygen sensing. Cell Metab. 1, 401–408 (2005). in mammalian development. RNA Biol. 17, 1214–1222 (2020).
90. Bell, E. L. et al. The Qo site of the mitochondrial complex III is required for the 126. Fagan, S. G., Helm, M. & Prehn, J. H. M. tRNA-derived fragments: a new class of
transduction of hypoxic signaling via reactive oxygen species production. J. Cell Biol. non-coding RNA with key roles in nervous system function and dysfunction.
177, 1029–1036 (2007). Prog. Neurobiol. 205, 102118 (2021).
91. Nawrot, B., Sochacka, E. & Duchler, M. tRNA structural and functional changes induced 127. Thompson, D. M. & Parker, R. Stressing out over tRNA cleavage. Cell 138, 215–219 (2009).
by oxidative stress. Cell Mol. Life Sci. 68, 4023–4032 (2011). 128. Maute, R. L. et al. tRNA-derived microRNA modulates proliferation and the DNA damage
92. Endres, L. et al. Alkbh8 regulates selenocysteine-protein expression to protect against response and is down-regulated in B cell lymphoma. Proc. Natl Acad. Sci. USA 110,
reactive oxygen species damage. PLoS ONE 10, e0131335 (2015). 1404–1409 (2013).
93. Chan, C. T. et al. Reprogramming of tRNA modifications controls the oxidative stress 129. Balatti, V. et al. TCL1 targeting miR-3676 is codeleted with tumor protein p53 in chronic
response by codon-biased translation of proteins. Nat. Commun. 3, 937 (2012). lymphocytic leukemia. Proc. Natl Acad. Sci. USA 112, 2169–2174 (2015).
94. Netzer, N. et al. Innate immune and chemically triggered oxidative stress modifies 130. Pekarsky, Y. et al. Dysregulation of a family of short noncoding RNAs, tsRNAs, in human
translational fidelity. Nature 462, 522–526 (2009). cancer. Proc. Natl Acad. Sci. USA 113, 5071–5076 (2016).

Nature Reviews Cancer


Review article

131. Wang, H. et al. The tRNA-derived fragment tRF-24-V29K9UV3IU functions as a miRNA-like 166. Olvedy, M. et al. A comprehensive repertoire of tRNA-derived fragments in prostate
RNA to prevent gastric cancer progression by inhibiting GPR78 expression. J. Oncol. cancer. Oncotarget 7, 24766–24777 (2016).
2022, 8777697 (2022). 167. Rich, A. & RajBhandary, U. L. Transfer RNA: molecular structure, sequence, and
132. Dong, D. D., Zhou, H. & Li, G. GPR78 promotes lung cancer cell migration and metastasis properties. Annu. Rev. Biochem. 45, 805–860 (1976).
by activation of Gαq-Rho GTPase pathway. BMB Rep. 49, 623–628 (2016). 168. Crothers, D. M., Seno, T. & Söll, D. G. Is there a discriminator site in transfer RNA?
133. Shi, Z. & Barna, M. Translating the genome in time and space: specialized ribosomes, Proc. Natl Acad. Sci. USA 69, 3063–3067 (1972).
RNA regulons, and RNA-binding proteins. Annu. Rev. Cell Dev. Biol. 31, 31–54 (2015). 169. Sprinzl, M., Horn, C., Brown, M., Ioudovitch, A. & Steinberg, S. Compilation of tRNA
134. Falconi, M. et al. A novel 3′-tRNAGlu-derived fragment acts as a tumor suppressor in breast sequences and sequences of tRNA genes. Nucleic Acids Res. 26, 148–153 (1998).
cancer by targeting nucleolin. FASEB J. 33, 13228–13240 (2019). 170. Shi, H. & Moore, P. B. The crystal structure of yeast phenylalanine tRNA at 1.93 Å
135. Guzzi, N. et al. Pseudouridine-modified tRNA fragments repress aberrant protein resolution: a classic structure revisited. RNA 6, 1091–1105 (2000).
synthesis and predict leukaemic progression in myelodysplastic syndrome. 171. Galli, G., Hofstetter, H. & Birnstiel, M. L. Two conserved sequence blocks within
Nat. Cell Biol. 24, 299–306 (2022). eukaryotic tRNA genes are major promoter elements. Nature 294, 626–631 (1981).
136. Guzzi, N. et al. Pseudouridylation of tRNA-derived fragments steers translational control 172. Hofstetter, H., Kressmann, A. & Birnstiel, M. L. A split promoter for a eucaryotic tRNA
in stem cells. Cell 173, 1204–1216.e26 (2018). gene. Cell 24, 573–585 (1981).
137. Pan, L. et al. Inflammatory cytokine-regulated tRNA-derived fragment tRF-21 suppresses 173. Takaku, H., Minagawa, A., Takagi, M. & Nashimoto, M. A candidate prostate cancer
pancreatic ductal adenocarcinoma progression. J. Clin. Invest. 131, e148130 (2021). susceptibility gene encodes tRNA 3′ processing endoribonuclease. Nucleic Acids Res.
138. Tong, L. et al. The tRNA-derived fragment-3017A promotes metastasis by inhibiting 31, 2272–2278 (2003).
NELL2 in human gastric cancer. Front. Oncol. 10, 570916 (2020). 174. Rideout, E. J., Marshall, L. & Grewal, S. S. Drosophila RNA polymerase III repressor
139. Zhang, F. et al. A 3′-tRNA-derived fragment enhances cell proliferation, migration and Maf1 controls body size and developmental timing by modulating tRNAiMet synthesis
invasion in gastric cancer by targeting FBXO47. Arch. Biochem. Biophys. 690, 108467 and systemic insulin signaling. Proc. Natl Acad. Sci. USA 109, 1139–1144 (2012).
(2020). 175. Chu, A. et al. Large-scale profiling of microRNAs for the cancer genome atlas.
140. Han, L. et al. A 5′-tRNA halve, tiRNA-Gly promotes cell proliferation and migration via Nucleic Acids Res. 44, e3 (2016).
binding to RBM17 and inducing alternative splicing in papillary thyroid cancer. J. Exp. 176. Cozen, A. E. et al. ARM-seq: AlkB-facilitated RNA methylation sequencing reveals
Clin. Cancer Res. 40, 222 (2021). a complex landscape of modified tRNA fragments. Nat. Methods 12, 879–884 (2015).
141. Liu, X. et al. A pro-metastatic tRNA fragment drives nucleolin oligomerization and 177. Zheng, G. et al. Efficient and quantitative high-throughput tRNA sequencing.
stabilization of its bound metabolic mRNAs. Mol. Cell 82, 2604–2617.e8 (2022). Nat. Methods 12, 835–837 (2015).
142. Lampson, B. L. et al. Rare codons regulate KRas oncogenesis. Curr. Biol. 23, 70–75 178. Smith, A. M., Abu-Shumays, R., Akeson, M. & Bernick, D. L. Capture, unfolding, and
(2013). detection of individual tRNA molecules using a nanopore device. Front. Bioeng.
143. Pershing, N. L. et al. Rare codons capacitate Kras-driven de novo tumorigenesis. Biotechnol. 3, 91 (2015).
J. Clin. Invest. 125, 222–233 (2015). 179. Thomas, N. K. et al. Direct nanopore sequencing of individual full length tRNA strands.
144. Li, S. & Counter, C. M. Signaling levels mold the RAS mutation tropism of urethane. eLife ACS Nano 15, 16642–16653 (2021).
10, e67172 (2021). 180. Hughes, L. A. et al. Copy number variation in tRNA isodecoder genes impairs mammalian
145. Ali, M. et al. Codon bias imposes a targetable limitation on KRAS-driven therapeutic development and balanced translation. Nat. Commun. 14, 2210 (2023).
resistance. Nat. Commun. 8, 15617 (2017). This is one of the first studies to systematically generate single and combinatorial
146. Keller, T. L. et al. Halofuginone and other febrifugine derivatives inhibit prolyl-tRNA deletions of related tRNA isodecoder genes and assess their consequences at an
synthetase. Nat. Chem. Biol. 8, 311–317 (2012). organismal level in mice.
147. Kim, Y. et al. Aminoacyl-tRNA synthetase inhibition activates a pathway that branches 181. Koukuntla, R., Ramsey, W. J., Young, W. B. & Link, C. J. U6 promoter-enhanced GlnUAG
from the canonical amino acid response in mammalian cells. Proc. Natl Acad. Sci. USA suppressor tRNA has higher suppression efficacy and can be stably expressed in 293
117, 8900–8911 (2020). cells. J. Gene Med. 15, 93–101 (2013).
148. Kim, S. H., Bae, S. & Song, M. Recent development of aminoacyl-tRNA synthetase
inhibitors for human diseases: a future perspective. Biomolecules 10, 1625 (2020). Acknowledgements
149. Zuko, A. et al. tRNA overexpression rescues peripheral neuropathy caused by mutations The authors thank the members of the Tavazoie laboratory for critically reading the
in tRNA synthetase. Science 373, 1161–1166 (2021). manuscript and for their helpful suggestions and comments. The authors apologize to all
150. Wang, J. et al. AAV-delivered suppressor tRNA overcomes a nonsense mutation in mice. the scientists whose work could not be cited owing to space limitations. S.F.T. and A.M.P.
Nature 604, 343–348 (2022). were supported by grants from the National Cancer Institute of the National Institutes of
151. Albers, S. et al. Engineered tRNAs suppress nonsense mutations in cells and in vivo. Health under award numbers R01CA257153, R35CA274446 and U54CA261701, as well as
Nature 618, 842–848 (2023). the Black Family Metastasis Center and the Breast Cancer Research Foundation. Molecular
152. Chang, J. C., Temple, G. F., Trecartin, R. F. & Kan, Y. W. Suppression of the nonsense graphics were performed with UCSF ChimeraX, developed by the Resource for Biocomputing,
mutation in homozygous β0 thalassaemia. Nature 281, 602–603 (1979). Visualization and Informatics at the University of California, San Francisco, with support
153. Winkle, M., El-Daly, S. M., Fabbri, M. & Calin, G. A. Noncoding RNA therapeutics — from National Institutes of Health R01-GM129325 and the Office of Cyber Infrastructure and
challenges and potential solutions. Nat. Rev. Drug Discov. 20, 629–651 (2021). Computational Biology, National Institute of Allergy and Infectious Diseases.
154. Papadimitriou, M. A. et al. tRNA-derived fragments (tRFs) in bladder cancer: increased
5′-tRF-LysCTT results in disease early progression and patients’ poor treatment outcome. Author contributions
Cancers 12, 3661 (2020). The authors contributed equally to all aspects of the article.
155. Wang, J. et al. Circulating tRNA-derived small RNAs (tsRNAs) signature for the diagnosis
and prognosis of breast cancer. NPJ Breast Cancer 7, 4 (2021). Competing interests
156. Wang, X. et al. Identification of tRNA-derived fragments expression profile in breast The authors declare no competing interests.
cancer tissues. Curr. Genomics 20, 199–213 (2019).
157. Sun, C. et al. tRNA-derived fragments as novel predictive biomarkers for Additional information
trastuzumab-resistant breast cancer. Cell. Physiol. Biochem. 49, 419–431 (2018). Peer review information Nature Reviews Cancer thanks Pierre Close and the other,
158. Veneziano, D. et al. Dysregulation of different classes of tRNA fragments in chronic anonymous, reviewer(s) for their contribution to the peer review of this work.
lymphocytic leukemia. Proc. Natl Acad. Sci. USA 116, 24252–24258 (2019).
159. Nientiedt, M. et al. Identification of aberrant tRNA-halves expression patterns in clear cell Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
renal cell carcinoma. Sci. Rep. 6, 37158 (2016). published maps and institutional affiliations.
160. Wu, Y. et al. 5′-tRF-GlyGCC: a tRNA-derived small RNA as a novel biomarker for colorectal
cancer diagnosis. Genome Med. 13, 20 (2021). Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this
161. Zhu, Y. et al. Comprehensive analysis of a tRNA-derived small RNA in colorectal cancer. article under a publishing agreement with the author(s) or other rightsholder(s); author
Front. Oncol. 11, 701440 (2021). self-archiving of the accepted manuscript version of this article is solely governed by the
162. Shao, Y. et al. tRF-Leu-CAG promotes cell proliferation and cell cycle in non-small cell terms of such publishing agreement and applicable law.
lung cancer. Chem. Biol. Drug Des. 90, 730–738 (2017).
163. Li, J., Cao, C., Fang, L. & Yu, W. Serum transfer RNA-derived fragment tRF-31-
79MP9P9NH57SD acts as a novel diagnostic biomarker for non-small cell lung cancer. Related links
J. Clin. Lab. Anal. https://doi.org/10.1002/jcla.24492 (2022). RCSB Protein Data Bank: https://www.rcsb.org/
164. Huang, Y. et al. Elucidating the role of serum tRF-31-U5YKFN8DYDZDD as a novel UCSF ChimeraX software: https://www.rbvi.ucsf.edu/chimerax
diagnostic biomarker in gastric cancer (GC). Front. Oncol. 11, 723753 (2021).
165. Panoutsopoulou, K. et al. tRNAGlyGCC-derived internal fragment (i-tRF-GlyGCC) in ovarian © Springer Nature Limited 2023
cancer treatment outcome and progression. Cancers 14, 24 (2021).

Nature Reviews Cancer

You might also like