ArticlePDF Available

Functional knockout of FUT8 in Chinese hamster ovary cells using CRISPR/Cas9 to produce a defucosylated antibody

Wiley
Engineering in Life Sciences
Authors:
  • Jecho Laboratories, Inc.

Abstract and Figures

We report the adaption of the new Crispr/Cas9 system to disrupt the gene encoding FUT8, an α-1,6-fucosyltransferase that directs fucose addition to derived antibody Fc region Asn 297, in Chinese hamster ovary (CHO) cells. Compared to previously reported homologous recombination (HR) or Zinc finger nucleases (ZFNs) applications in CHO cells, Crispr/Cas9 demonstrated higher targeting efficiency and easier customization. FUT8 disruptive clones (FUT8-/-) were obtained within 3 weeks at indel frequencies ranged from 9% to 25%, which could be enhanced to 52% with Lens culinaris agglutinin (LCA) selection. Based on the lectin blot method, the derived FUT8-/- clone had the ability producing defucosylated therapeutic monoclonal antibody with no detrimental effects on cell growth, viability, or product quality. The clone had the potential of industrial application for therapeutic antibodies manufacturing. We have demonstrated functionally that a gene related to product synthesis could be mutated using Crispr/Cas9 technology and consequently the glycan profile of expressed monoclonal antibody was alternated. We believe that with its robustness and effectiveness, Crispr/Cas9 method can be widely applicable in cell line development leading to higher productivity and better quality of monoclonal antibodies and other biological therapeutics.This article is protected by copyright. All rights reserved
Content may be subject to copyright.
Eng. Life Sci. 2015,15, 660–666 www.els-journal.com
Tao Sun1,2
Chaodong Li1,2
Lei Han1,2
Hua Jiang3
Yueqing Xie3
Baohong Zhang1,2
Xiuping Qian1,2
Huili Lu1,2
Jianwei Zhu1,2, 3
1School of Pharmacy, Shanghai
Jiao Tong University, Shanghai,
China
2Engineering Research Center of
Cell & Therapeutic Antibody,
Ministry Of Education (MOE),
Shanghai, China
3Jecho Laboratories, Inc.,
Frederick, MD, USA
Short Communication
Functional knockout of FUT8 in Chinese
hamster ovary cells using CRISPR/Cas9 to
produce a defucosylated antibody
We report the adaptation of the new CRISPR/Cas9 (clustered regularly interspaced
short palindromic repeat/CRISPR-associated protein 9) system to disrupt the gene
encoding fucosyltransferase 8 (FUT8), an α1,6-fucosyltransferase that directs fu-
cose addition to derived antibody Fc region asparagine 297, in Chinese hamster
ovary (CHO) cells. Compared to previously reported homologous recombination
or zinc-finger nucleases (ZFNs) applications in CHO cells, CRISPR/Cas9 demon-
strated higher targeting efficiency and easier customization. FUT8 disruptive clones
(FUT8/) were obtained within 3 weeks at indel frequencies ranging from 9 to
25%, which could be enhanced to 52% with Lens culinaris agglutinin (LCA) selec-
tion. Based on the lectin blot method, the derived FUT8/clonehadtheabilityto
produce defucosylated therapeutic mAb with no detrimental effects on cell growth,
viability, or product quality. The clone had the potential of industrial application for
therapeutic antibodies manufacturing. We have demonstrated functionally that a
gene related to product synthesis could be mutated using CRISPR/Cas9 technology,
and consequently the glycan profile of expressed mAb was alternated. We believe
that with its robustness and effectiveness, CRISPR/Cas9 can be widely applicable in
cell line development leading to higher productivity and better quality of mAbs and
other biological therapeutics.
Keywords: Antibody / Antibody-dependent cellular cytotoxicity / CHO-K1 cells / CRISPR/Cas9
/ Fucose
Additional supporting information may be found in the online version of this article at
the publisher’s web-site
Received: October 23, 2014; revised: May 29, 2015; accepted: June 15, 2015
DOI: 10.1002/elsc.201400218
1 Introduction
mAb therapeutics is increasing rapidly in treatment of various
diseases [1]. There are more than 30 antibody therapeutics ap-
proved by regulatory agencies for the worldwide market, and
many more mAb candidates are in clinical trials [2]. Antibod-
ies execute their function by destroying targets through two
Correspondence: Jianwei Zhu (jianweiz@sjtu.edu.cn), School of
Pharmacy, Shanghai Jiao Tong University, Minhang 200240, Shang-
hai, China
Abbreviations: ADCC, antibody-dependent cellular cytotoxicity; CHO,
Chinese hamster ovary; CRISPR/Cas9, clustered regularly interspaced
short palindromic repeat/CRISPR-associated protein 9; FUT8,fucosyl-
transferase 8 (α1,6-fucosyltransferase); LCA,Lens culinaris agglutinin;
sgRNA, single guide RNA; TALEN , transcription activator-like effector
nuclease; ZFN, zinc-finger nuclease
pathways: complement-dependent cytotoxicity and antibody-
dependent cellular cytotoxicity (ADCC) [3]. ADCC is mediated
by antibody Fc region binding to lymphocyte receptors [4]. N-
oligosaccharide is present on asparagine 297 of IgG heavy chain
Fc region, whereas fucose is commonly found in the oligosac-
charide structure, which could hinder Fc binding to lymphocyte
receptors [5]. Compared to fucosylated antibodies derived from
wild-type Chinese hamster ovary (CHO) cells, defucosylated
antibodies achieved 100-fold higher ADCC activity in vitro [6].
Therefore, antibody defucosylation is considered to be a power-
ful approach to enhance ADCC activity.
The majority of commercial antibodies are produced from
CHO cells, with advantages of posttranslational modifications,
Additional correspondence: Huili Lu (roadeer@sjtu.edu.cn), School of
Pharmacy, Shanghai Jiao Tong University, Minhang 200240, Shanghai
China
660 C2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.els-journal.com Eng. Life Sci. 2015,15, 660–666 www.biotecvisions.com
such as glycosylation, phosphorylation, acetylation, etc [7].
However, it is notable that core fucosylation commonly found
in mAbs derived from CHO cells would hinder antibodies bind-
ing to their receptors [8]. In CHO cells, fucosyltransferase 8
(FUT8) is an α1,6-fucosyltransferase that directs fucose addition
to asparagine-linked N-acetylglucosamine moieties, a common
feature of N-linked glycan core structures [9]. So far, the effect of
disrupting FUT8 gene to enhance ADCC has been reported by
several groups using different approaches, such as homologous
recombination [10] or zinc-finger nucleases (ZFNs) [11], both
achieved increased ADCC activity (approximately 100-fold).
Traditionally, homologous recombination was used for
genome editing in cell lines and animal models, though it was
less efficient and was a time-consuming process due to technical
complexity [12]. This was particularly true in the case of FUT8
disruption, as approximately 120 000 clones were screened to
obtain only three FUT8/clones with a normal growth pro-
file [13]. To improve the efficiency of disrupting a coding gene,
technologies such as ZFNs and transcription activator-like ef-
fector nucleases (TALENs) were developed and reported [14].
Compared to traditional homologous recombination, ZFNs and
TALENs demonstrated higher genome editing efficiency [15].
These nucleases could introduce double-strand breaks, which
would be repaired by one of the following two major mecha-
nisms: the error-prone nonhomologous end-joining or the high-
fidelity homology-directed repair pathway [16]. Nonhomolo-
gous end-joining can be harnessed to mediate gene functional
disruptions, as indels occurring in a coding region could lead to
frameshift mutations and protein function loss [17]. ZFNs and
TALENs were extremely expensive and difficult to design, which
limited their widespread use. More recently, CRISPR (clustered
regularly interspaced short palindromic repeat)/Cas9 (CRISPR-
associated protein 9) system, which was derived from the mi-
crobial adaptive immune system [18], has been developed as an
efficient genome editing tool and applied to diverse species such
as plants, animals, bacteria, and yeast [19], as well as in CHO
cells [20]. This system required only Cas9 and single guide RNA
(sgRNA) [21], giving it several advantages including ease of cus-
tomization, higher targeting efficiency, and ability to facilitate
multiplex genome editing [22].
Here,wereportedanewproceduretodisruptFUT8
gene in CHO-K1 cell line using CRISPR/Cas9 system. Lens
culinaris agglutinin (LCA) based phenotypic screen was used
to enrich clones with exclusively doubly modified alleles. mAbs
produced in the CHO FUT8/cells were defucosylated.Further,
we demonstrated that the gene modification had no detrimental
effects on cell growth, viability, or product quality.
2 Materials and methods
2.1 Construction of FUT8 disruptive vector
The exon 10 of FUT8 (GeneBank ID: 100751648) was selected
as the target site for mutagenesis because it encoded for the cat-
alytic site of the enzyme. Five sgRNAs were designed, based
on previously described design rules [19] to obtain satisfied
indel frequencies (Supporting information Tables S1–S3 and
Supporting information, Fig. S1). Exon 10 and adjacent re-
gion (685 bp) was amplified by PCR using primers FUT8-F
(5’-CTGTTGATTCCAGGTTCCCATATA-3) and FUT8-R (5-
TTGA ATGATGACTGCTA GTGATGCT -3). The plasmid pX330
was kindly provided by Dr. Zhang Feng (MIT) [19] and digested
with BbsI (New England Biolabs, MA, USA); a pair of annealed
sgRNA was ligated to the linear plasmid. All cloning steps were
confirmed by DNA sequencing at Invitrogen (Shanghai, China),
derived plasmids were addressed as pX330-sgRNA15.
2.2 Analysis of Cas9 activity in CHO-K1 cells
Cells were incubated at 37C with 5% CO2and transfected using
Lipofectamine 2000 reagent (Invitrogen, CA, USA). Forty-eight
hours posttransfection cells were harvested and genome DNA
was extracted using Genome DNA Extraction Kit (Axygen, CA,
USA). Approximately 200 ng PCR product of FUT8 exon 10 re-
gion (685 bp) was annealed using the following program: heated
to 95C for 3 min and then ramped down to 25Cat5
C/min.
Annealed PCR fragment was treated with 0.3 μL T7 endonucle-
ase at 37C for 45 min, followed by analysis on 1% agarose gel.
Related indels were stained with ethidium bromide and detected
under gel imaging systems (Tanon, Shanghai, China). Percent-
age of indel frequency was calculated using ImageJ software as
previously described by Hwang et al. [23].
2.3 Generation of CHO FUT8/clones
Cells were exposed to 100 μg/mL LCA (Vector Labs, CA, USA)
2 days posttransfection. After 6-day growth in the presence of
100 μg/mL LCA, clones morphologically similar to the wild-type
CHO-K1 were expanded. Limiting dilution was performed to
screen doubly modified alleles’ clones. When clones reached an
80% confluence, genome DNA extraction and T7 endonuclease
digestion were performed as described above. To identify both
alleles’ modification in LCA-resistant clones, PCR product from
each T7E1-positive clone was TA-cloned using TA cloning kit
(Takara, Tokyo, Japan). At least 10 colonies were picked from
each transformation for DNA sequencing as CHO cells were
aneuploids.
2.4 LCA reactivity and genetic stability analysis
FITC-LCA binding assay was performed to test FUT8/clones
phenotypic profiles. Briefly described, cells were seeded in 12-
well plates and incubated in the presence of 20 μg/mL FITC-
LCA (Vector Labs) for 30 min. Cells were washed three times
and visualized under fluorescence microscope (Olympus, Tokyo,
Japan). An LCA-binding test based on flow cytometer analysis
was also performed as previously described [13]. Clones were
cultured for 10 passages to confirm the genetic stability of the
modified clones, both alleles’ DNA sequencing was performed
as described above.
2.5 Anti-CD20 mAb expression and purification
Chimeric anti-CD20 mAb (Rituximab, drug bank: DB00073)
was chosen as a model antibody [24]. The heavy chain and light
chain were separately subcloned into pcDNA3.1 (Invitrogen)
C2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 661
www.els-journal.com Eng. Life Sci. 2015,15, 660–666 www.biotecvisions.com
as expression vectors (heavy chain and light chain sequences
are shown in Supporting information). All cloning steps were
confirmed by DNA sequencing. Stable transfected cell pools of
CHO-K1 cells and FUT8/clones were selected via 800 μg/mL
G418 pressure for 10 days. Serum-free adaptationwas performed
to eliminate the influence of FBS presence. Cell count and via-
bility were determined by Trypan blue method. AKTA purifier
system (GE Healthcare, CT, USA) was used to purify antibodies
from culture supernatants as previously described [25].
2.6 SDS-PAGE, western blot, and lectin blot
Protein fractions from protein A chromatograph were analyzed
by 10% SDS-PAGE to confirm that purified glycoproteins were
antibodies. Protein loading was normalized to 500 ng/well for
SDS-PAGE, 50 ng/well for lectin blot, and 5 ng/well for western
blot. For lectin blot, the membrane was incubated with 10 μg/mL
FITC-LCA for 30 min in the dark, and the green fluorescence
signal was scanned by multiple function laser scanner (Typhoon
FLA 9500, GE Healthcare).
3 Results
3.1 FUT8 modification generated by CRISPR/Cas9
The exon 10 encoded for the catalytic site inverting substrate
GDP-fucose, therefore was chosen as the target for CRISPR/Cas9
editing. Plasmid containing a chimeric guide RNA structure
had higher genome editing efficiency as compared to plasmid
containing a crRNA-tracrRNA guide RNA [19], hence plasmid
pX330 containing human U6 polymerase promoter generat-
ing chimeric sgRNA was selected to perform gene perturba-
tion, which was confirmed to be effective in multiple species
as well as Chinese hamster [20]. Two days posttransfection, T7
endonuclease digestion tests were performed in cells transfected
with diverse pX330-sgRNAs. Indels were stained with ethidium
bromide and visualized under DNA gel imaging system, indel
frequencies (%) were analyzed using ImageJ software with the
value ranging from 9 to 24% as indicated in Fig. 1A. No indels
were visualized without pX330-sgRNA treatment.
3.2 FUT8 functional disruptive clones generated by
limiting dilution
To produce defucosylation mAbs in CRISPR/Cas9 derived
clones, both alleles of FUT8 gene should be disrupted. LCA is a
plant lectin with high specificity to bind IgG N-glycan core fu-
cose, which can lead to cell death [26]. CHO FUT8/clones are
able to survive in the presence of LCA as cells fail to synthesize fu-
cosylated glycoproteins. Inour experiment, cells were transferred
to 12-well plates 2 days posttransfection and LCA was added to
each well at a final concentration of 100 μg/mL (day 2). Eighteen
hours later, most cells became round and got detached from the
plate, the phenotype similar to wild-type clones appeared 2 days
after LCA treatment (day 4, Fig. 1B). Cells were LCA selected for
6 days and expanded after fresh medium replacement (day 8).
As shown in Fig. 1A, indel frequencies of LCA-resistant clones
were ranging from 24 to 52%, indicating that LCA treatment en-
riched FUT8/clones. No LCA-resistant clones were detected
in the absence of pX330-sgRNA transfection within 6 days LCA
exposure.
LCA-selected clones derived from pX330-sgRNA1 transfec-
tion were subcloned by limiting dilution. Three clones, named
D5, C9, and C10 showing positive T7E1 digestion activity were
chosen for further research. Exon 10 region was amplified by
PCR from these clones, both alleles’ genotypes of clone D5, C9,
and C10 are listed in Fig. 1C. These modifications were consid-
ered to lead FUT8 truncation and frameshift mutations in these
clones. C9 clone was morphologically different from wild-type
CHO cells (Supporting information, Fig. S2), which indicated
that other alternations in cell metabolism might occur besides
the FUT8 disruption. C10 clone was successfully adapted to
serum-free medium and had the best cell growth among the
three clones, therefore was selected for further studies.
There was no available commercial antibody that could dis-
tinguish FUT8 exon 10 functional disruption. We therefore tested
the FUT8 activity by LCA-resistance phenotype method. It was
confirmed by fluorescence microscopy images and flow cytome-
ter analysis that these clones lacked the ability to bind FITC-LCA.
As indicated in Fig. 2A and B and Supporting information Fig. S3,
FITC-LCA could bind to wild-type CHO-K1 cells but not the
FUT8/clones. Genetic stability of C10 clone was also evalu-
ated for 10 passages. No alternation was found in genotype or
LCA-resistance ability compared to the parental clone (data not
shown).
3.3 Antibodies expression and lectin blot based
fucosylation detection
Prior to express an antibody in the cells, growth profiles of C10
clone and wild-type CHO-K1 cells were compared side by side
in 125-mL shaker flasks with 20 mL working-volume medium.
During 8 days incubation, both CHO-K1 cells and C10 clone
cell densities reached greater than 6 ×106cells/mL and main-
tained viability above 90%. As shown in Fig. 2C and D, C10
clone had comparable viable cell counts and culture viability
compared to wild-type CHO-K1 cells throughout three parallel
experiments.
Cell pools stably expressing anti-CD20 antibody were derived
from both wild-type CHO-K1 cells and C10 clone using G418 se-
lection. Each cell pool was cultured in a 1000-mL shake flask con-
taining 300 mL CD-CHO medium supplemented with HT and
glutamine, supernatants were collected when cell viabilities were
below 30%. Proteins of interest in the supernatants were purified
by AKTA purifier system and analyzed by SDS-PAGE and western
blot (Fig. 3A and B). Roughly equivalent quantity of purified IgGs
were obtained from C10 clone and CHO-K1 cells pools under
the same cell incubation condition based on bicinchoninic acid
assays, with expression levels around 1.2 mg/L. FITC-LCA lectin
blot was chosen as an alternative method to examine whether
antibody derived from the C10 clone contained fucose [27].
As expected, antibody derived from C10 clone showed no visi-
ble fluorescence compared to antibody from parental CHO-K1
cells under both nonreducing and reducing conditions in three
662 C2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.els-journal.com Eng. Life Sci. 2015,15, 660–666 www.biotecvisions.com
Figure 1. Genome editing in CHO-K1 cells by CRISPR/Cas9. (A) T7E1 digestion for Cas9 activity in CHO cells. Two hundred nanograms
PCR product was digested and loaded for each lane. Left panel, genomic DNA was extracted from CHO-K1 cells 2 days after transfected
with different pX330-sgRNAs. Right panel, genomic DNA was extracted from expanded clones by LCA selection. Note that indel frequency
was not completely equivalent to disruption efficiency. (B) Wild-type CHO-K1 cells were round and detached from the plate when treated
with 100 μg/mL LCA for 18 h, while pX330-sgRNAs transfected clones with wild-type morphology appeared under the same condition. LCA
precipitate was visible. (C) Sanger sequences of FUT8 exon 10 mutated alleles obtained from three CHO FUT8/clones. Alleles were
designated as A and B. The gap shown in the reference sequences has been included for better representation of alignment results. All
modifications could lead to FUT8 frameshift mutations.
parallel experiments (Fig. 3C). Weconcluded that CHO FUT8/
clone generated by CRISPR/Cas9 was able to express mAb with
significantly reduced fucosylation.
4 Discussion
In this report, we presented successful FUT8-functional disrup-
tion in CHO-K1 cells using CRISPR/Cas9 system. Normally, the
20–30 bp region downstream of the start codon ATG was chosen
as the target [28], while catalytic site of FUT8 protein was chosen
for successful gene perturbation in CHO-K1 cells as previously
reported [8]. Off-target mutagenesis was always an undesirable
effect induced by Cas9 and hard to avoid, enhancement of Cas9
targeting specificity seemed to be promising as many researchers
had fruitful achievements of improved Cas9 target recognition
fidelity [29, 30]. A successful gene disruption in COSMC and
FUT8 with high indel frequency was reported, in addition to
gene disruption results that showed consistency as reported by
Ronda et al. [20], we were able to demonstrate the final product
without fucose.
C2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 663
www.els-journal.com Eng. Life Sci. 2015,15, 660–666 www.biotecvisions.com
Figure 2. LCA reactivity and cell growth of CHO FUT8/C10 clone. (A) FITC-LCA failed to bind C10 clone membrane glycoproteins. Under
panels, cells were treated with 20 μg/mL FITC-LCA, fluorescence microscopy images were subsequently acquired at a green fluorescent
protein channel. (B) FACS paragraphs of FITC-LCA stained cells. x-axis, fluorescence intensity; y-axis, cell counts. NC refers to negative
control wild-type CHO-K1 cells without FITC-LCA treatment. For details see Supporting information Fig. S3. Cell growth (C) and viability (D)
of serum-free adapted C10 clone and CHO-K1 cells. Cell density and cell viability were examined in 125-mL shaker flasks (20 mL working
volume) during 8 days incubation. Error bars represent the SD from three parallel experiments.
LCA selection was conducted to obtain stable FUT8/clones
within a short duration. Another gene related to core fucose
synthesis pathway named GDP-fucose transporter was also dis-
rupted in our lab (data not shown), which was responsible for
transporting GDP-fucose into Golgi apparatus [31]. GDP-fucose
transporter modified clones showed no resistance to LCA, indi-
cated that other GDP-fucose transporting pathways could play
complementary roles.
A new host cell line to produce completely defucosylated an-
tibodies, needs to be characterized in terms of growth, viability,
and antibody productivity [32]. In this study, we compared the
C10 clone and wild-type CHO-K1 cells in a shake flask sys-
tem. Cell cultures were maintained for 8 days before significant
clumping occurred. For further comparison, feb-batch cultiva-
tion in a bioreactor system would be preferable. Our data showed
CHO FUT8/C10 clone had comparable viable cell counts and
viability throughout three parallel experiments compared to the
wild-type cell line.
To detect the presence of fucose in the mAb glycan structure,
we employed a simple and quick FITC-LCA lectin blot method,
whereas other approaches such as MALDI-TOLF-MS and high
pH anion exchange chromatography with pulsed amperometric
detection could be utilized [33]. FITC-LCA showed very high
specificity binding to α1,6-fucose and was a widely accepted
approach to confirm N-glycan fucosylation of glycoproteins [27,
34]. As expected, antibody produced in C10 clone showed no
detectable fluorescence compared to antibody from the parental
CHO-K1 cells under both nonreducing and reducing conditions,
indicating that the C10 clone produced defucosylated antibodies.
In summary, our derived CHO FUT8/clone had the ability
to produce defucosylated antibody with no detrimental effects
in cell growth, viability, and the potential for industrial utiliza-
tion. We demonstrated CRISPR/Cas9 application in CHO-K1
cells genome editing as well as its effect on product glycosyla-
tion alternation, which would be potentially useful in indus-
trial purposes [35, 36]. To develop the potential of CRISPR/Cas9
664 C2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.els-journal.com Eng. Life Sci. 2015,15, 660–666 www.biotecvisions.com
Figure 3. Biochemical analysis of IgG purified from C10 clone and CHO-K1 cells transfected with anti-CD20 mAb expression vectors. (A)
SDS-PAGE and Coomassie brilliant blue staining of the purified IgG. (B) Western blot of the purified IgG. Anti-CD20 mAbs were detected
by HRP-labeled goat antihuman Fc antibody and ECL. (C) The PVDF membrane was incubated with 10 μg/mL FITC-LCA, and subsequently
detected using multifunctional laser scanner after washing three times under a 488 nm channel. Note that proteins were normalized to
500 ng/well for SDS-PAGE, 50 ng/well for lectin blot, and 5 ng/well for western blot.
application in bioprocess, other genes related to recombinant
protein productivity or quality enhancement are currently be-
ing investigated in our labs. We believe that as a robust genome
editing tool, CRISPR/Cas9 system can be widely applicable in
cell line development for manufacturing high-quality mAbs or
other biological therapeutics.
We thank F. Zhang of the Broad Institute of MIT and Harvard for
kindly providing us with the Cas9 expression vector. The research
was supported by Chinese Natural Science Foundation #81473127.
The authors have declared no conflict of interest.
5 References
[1] Deb, R., Chakraborty, S., Veeregowda, B., Verma, A. K. et al.,
Monoclonal antibody and its use in the diagnosis of livestock
diseases. Adv. Biosci. Biotechnol. 2013, 4, 50–62.
[2] Elvin, J. G., Couston, R. G., vander Walle, C. F., Therapeutic
antibodies: Market considerations, disease targets and biopro-
cessing. Int. J. Pharmaceutics 2013, 440, 83–98.
[3] Cartron, G., Dacheux, L., Salles, G., Solal-Celigny, P. et al.,
Therapeutic activity of humanized anti-CD20 monoclonal an-
tibody and polymorphism in IgG Fc receptor FcγRIIIa gene.
Blood 2002, 99, 754–758.
[4] Clynes, R. A., Towers, T. L., Presta, L. G., Ravetch, J. V., In-
hibitory Fc receptors modulate in vivo cytoxicity against tumor
targets. Nat. Med. 2000, 6, 443–446.
[5] Rademacher, T., Homans, S., Parekh, R., Dwek, R., Im-
munoglobulin G as a glycoprotein. Biochem. Soc. Symp. 1985,
51, 131–148.
[6] Peipp, M., vanBueren, J. J. L., Schneider-Merck, T., Bleeker, W.
W. et al., Antibody fucosylation differentially impacts cytotox-
icity mediated by NK and PMN effector cells. Blood 2008, 112,
2390–2399.
[7] Zhu, J., Mammalian cell protein expression for biopharmaceu-
tical production. Biotechnol. Adv. 2012, 30, 1158–1170.
[8] Kanda, Y., Yamane-Ohnuki, N., Sakai, N., Yamano, K. et al.,
Comparison of cell lines for stable production of fucose-
negative antibodies with enhanced ADCC. Biotechnol. Bioeng.
2006, 94, 680–688.
[9] Becker, D. J., Lowe, J. B., Fucose: Biosynthesis and bi-
ological function in mammals. Glycobiology 2003, 13,
41R–53R.
[10] Shinkawa, T., Nakamura, K., Yamane, N., Shoji-Hosaka, E.
et al., The absence of fucose but not the presence of galactose or
bisecting N-acetylglucosamine of human IgG1 complex-type
oligosaccharides shows the critical role of enhancing antibody-
dependent cellular cytotoxicity. J. Biol. Chem. 2003, 278, 3466–
3473.
[11] Malphettes,L.,Freyvert,Y.,Chang,J.,Liu,P.Q.etal.,Highly
efficient deletion of FUT8 in CHO cell lines using zinc-finger
nucleases yields cells that produce completely nonfucosylated
antibodies. Biotechnol. Bioeng. 2010, 106, 774–783.
[12] Zwaka, T. P., Thomson, J. A., Homologous recombination in
human embryonic stem cells. Nat. Biotechnol. 2003, 21, 319–
321.
[13] Yamane-Ohnuki, N., Kinoshita, S., Inoue-Urakubo, M.,
Kusunoki, M. et al., Establishment of FUT8 knockout Chi-
nese hamster ovary cells: An ideal host cell line for producing
completely defucosylated antibodies with enhanced antibody-
dependent cellular cytotoxicity. Biotechnol. Bioeng. 2004, 87,
614–622.
[14] Mussolino, C., Morbitzer, R., L¨
utge, F., Dannemann, N. et al.,
A novel TALE nuclease scaffold enables high genome editing
activity in combination with low toxicity. Nucleic Acids Res.
2011, 39, 9283–9293.
[15] Walsh, R. M., Hochedlinger, K., A variant CRISPR-Cas9 system
adds versatility to genome engineering. Proc. Natl. Acad. Sci.
2013, 110, 15514–15515.
[16] Urnov, F. D., Rebar, E. J., Holmes, M. C., Zhang, H. S. et al.,
Genome editing with engineered zinc finger nucleases. Nat.
Rev. Genet. 2010, 11, 636–646.
[17] Perez, E. E., Wang, J., Miller, J. C., Jouvenot, Y. et al., Es-
tablishment of HIV-1 resistance in CD4+Tcellsbygenome
editing using zinc-finger nucleases. Nat. Biotechnol. 2008, 26,
808–816.
[18] Barrangou, R., Fremaux, C., Deveau, H., Richards, M.
et al., CRISPR provides acquired resistance against viruses in
prokaryotes. Science 2007, 315, 1709–1712.
C2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 665
www.els-journal.com Eng. Life Sci. 2015,15, 660–666 www.biotecvisions.com
[19] Cong,L., Ran, F. A., Cox, D., Lin, S. et al., Multiplex genome en-
gineering using CRISPR/Cas systems. Science 2013, 339, 819–
823.
[20] Ronda, C., Pedersen, L. E., Hansen, H. G., Kallehauge, T. B.
et al., Accelerating genome editing in CHO cells using CRISPR
Cas9 and CRISPy, a web-based target finding tool. Biotechnol.
Bioeng. 2014, 111, 1604–1616.
[21] Mali, P., Yang, L., Esvelt, K. M., Aach, J. et al., RNA-guided
human genome engineering via Cas9. Science 2013, 339, 823–
826.
[22] Ran,F.A.,Hsu,P.D.,Wright,J.,Agarwala,V.etal.,Genome
engineering using the CRISPR-Cas9 system. Nat. Protoc. 2013,
8, 2281–2308.
[23] Hwang,W.Y.,Fu,Y.,Reyon,D.,Maeder,M.L.etal.,Efcient
genome editing in zebrafish using a CRISPR-Cas system. Nat.
Biotechnol. 2013, 31, 227–229.
[24] Lim, S. H., Beers, S. A., French, R. R., Johnson, P. W. et al.,
Anti-CD20 monoclonal antibodies: Historical and future per-
spectives. Haematologica 2010, 95, 135–143.
[25] Hober, S., Nord, K., Linhult, M., Protein A chromatogra-
phy for antibody purification. J. Chromatogr. B 2007, 848,
40–47.
[26] Miyoshi, E., Moriwaki, K., Terao, N., Tan, C.-C. et al., Fucosy-
lation is a promising target for cancer diagnosis and therapy.
Biomolecules 2012, 2, 34–45.
[27] Wang, X., Gu, J., Ihara, H., Miyoshi, E. et al., Core fucosylation
regulates epidermal growth factor receptor-mediated intracel-
lular signaling. J. Biol. Chem. 2006, 281, 2572–2577.
[28] Hsu, P. D., Lander, E. S., Zhang, F., Development and applica-
tions of CRISPR-Cas9 for genome engineering. Cell 2014, 157,
1262–1278.
[29] Fu,Y.,Foden,J.A.,Khayter,C.,Maeder,M.L.etal.,High-
frequency off-target mutagenesis induced by CRISPR-Cas nu-
cleases in human cells. Nat. Biotechnol. 2013, 31, 822–826.
[30] Ran,F.,Hsu,P.D.,Lin,C.-Y.,Gootenberg,J.S.etal.,Double
nicking by RNA-guided CRISPR Cas9 for enhanced genome
editing specificity. Cell 2013, 154, 1380–1389.
[31] L¨
ubke, T., Marquardt, T., Etzioni, A., Hartmann, E. et al., Com-
plementation cloning identifies CDG-IIc, a new type of con-
genital disorders of glycosylation, as a GDP-fucose transporter
deficiency. Nat. Genet. 2001, 28, 73–76.
[32] Rajendra, Y., Kiseljak, D., Baldi, L., Hacker, D. L. et al., A
simple high-yielding process for transient gene expression in
CHO cells. J. Biotechnol. 2011, 153, 22–26.
[33] Coulier, L., Zha, Y., Bas, R., Punt, P. J., Analysis of oligosac-
charides in lignocellulosic biomass hydrolysates by high-
performance anion-exchange chromatography coupled with
mass spectrometry (HPAEC–MS). Bioresour. Technol. 2013,
133, 221–231.
[34] Qiu, Y., Patwa, T. H., Xu, L., Shedden, K. et al., Plasma glyco-
protein profiling for colorectal cancer biomarker identification
bylectinglycoarrayandlectinblot.J. Proteome Res. 2008, 7,
1693–1703.
[35] Tsai,S.Q.,Wyvekens,N.,Khayter,C.,Foden,J.A.etal.,Dimeric
CRISPR RNA-guided FokI nucleasesfor hig hly specific genome
editing. Nat. Biotechnol. 2014, 32, 569–576.
[36] Zhu, J., Mammalian cell protein expression for biopharmaceu-
tical production. Biotechnol. Adv. 2012, 30, 1158–1170.
666 C2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
... This method significantly increased the targeting efficiency, elevating it from the previously reported range of 10~30% to an impressive 72%. This enhancement also ensured the precise integration of exogenous genes at designated sites [22]. Stable cell line development is based on productivity and stability. ...
... The adherent cells were cultured in Ham's F-12 K medium with 10% FBS at 37 • C in a 5% CO 2 incubator [22] (Sun 2015). Suspension cells were cultivated in a shaker using CD-CHO medium, supplemented with 8 mM L-glutamax, at 110 rpm, in the humified environment at 37 • C with 5% CO 2 . ...
... The adherent cells were cultured in Ham's F-12 K medium with 10% FBS at 37 • C in a 5% CO 2 incubator [22] (Sun 2015). Suspension cells were cultivated in a shaker using CD-CHO medium, supplemented with 8 mM L-glutamax, at 110 rpm, in the humified environment at 37 • C with 5% CO 2 . ...
Article
Full-text available
Chinese hamster ovary (CHO) cells are a popular choice in biopharmaceuticals because of their beneficial traits, including high-density suspension culture, safety, and exogenously produced proteins that closely resemble natural proteins. Nevertheless, a decline in the expression of exogenous proteins is noted as culture time progresses. This is a consequence of foreign gene recombination into chromosomes by random integration. The current investigation employs CRISPR-Cas9 technology to integrate foreign genes into a particular chromosomal location for sustained expression. Results demonstrate the successful integration of enhanced green fluorescent protein (EGFP) and human serum albumin (HSA) near base 434814407 on chromosome NC_048595.1 of CHO-K1 cells. Over 60 successive passages, monoclonal cell lines were produced that consistently expressed all relevant external proteins without discernible variation in expression levels. In conclusion, the CHO-K1 cell locus, NC_048595.1, proves an advantageous locus for stable exogenous protein expression. This study provides a viable approach to establishing a CHO cell line capable of enduring reliable exogenous protein expression.
... This review sums up the CRISPR-Cas9 system applications in CHO cell line engineering ( Figure 1). [17][18][19][20][21][22][23]; (B) targeted gene knockin [24][25][26][27]; (C) CRISPR activation [28][29][30]; (D) CRISPR interference [31][32][33]; (E) DNA base editing [34]; (F) targeted mRNA knockdown [35,36]. ...
... The reason for the widespread and rapid adoption of CRISPR technology lies in its simple RNA-DNA interactions for targeting. In contrast to earlier ZFNs and TALENs, which use protein-DNA interactions, complex and labor-intensive protein design is no longer needed, because RGENs use simple Watson- [17][18][19][20][21][22][23]; (B) targeted gene knockin [24][25][26][27]; (C) CRISPR activation [28][29][30]; (D) CRISPR interference [31][32][33]; (E) DNA base editing [34]; (F) targeted mRNA knockdown [35,36]. ZFNs, TALENs, and CRISPR-Cas9. ...
... The first report about CRISPR-Cas9-mediated KO of the Fut8 gene in CHO cells [57] was soon followed by a demonstration of the industrial potential of CRISPR-Cas9 technology to produce afucosylated antibodies [17]. The authors targeted exon 10 encoding for the catalytic site of the FUT8 enzyme in the CHO-K1 host cell line. ...
Article
Full-text available
The Chinese hamster ovary (CHO) cell line is a well-established platform for the production of biopharmaceuticals due to its ability to express complex therapeutic proteins with human-like glycopatterns in high amounts. The advent of CRISPR technology has opened up new avenues for the engineering of CHO cell lines for improved protein production and enhanced product quality. This review summarizes recent advances in the application of CRISPR technology for CHO cell line engineering with a particular focus on glycosylation modulation, productivity enhancement , tackling adventitious agents, elimination of problematic host cell proteins, development of antibiotic-free selection systems, site-specific transgene integration, and CRISPR-mediated gene activation and repression. The review highlights the potential of CRISPR technology in CHO cell line genome editing and epigenetic engineering for the more efficient and cost-effective development of biopharmaceuticals while ensuring the safety and quality of the final product.
... One approach involves cell line engineering using CRISPR-Cas9based gene editing to reduce or eliminate expression of genes encoding problematic HCPs [138]. By targeting specific genes responsible for HCP production, such as lipoprotein lipase (LPL), CHO cell lines were engineered to produce mAbs with substantially reduced amounts of this problematic HCP [139,140]. CRISPR-Cas9 technology, which has high target selectivity and is cost effective, is a promising technique for eliminating problematic HCPs that are not essential to the growth, survival, or cellular production of the protein therapeutic. Other techniques focus on downstream processing and purification steps. ...
Article
Full-text available
Monoclonal antibodies (mAbs) have transformed therapeutic strategies for various diseases. Their high specificity to target antigens makes them ideal therapeutic agents for certain diseases. However, a challenge to their application in clinical practice is their potential risk to induce unwanted immune response, termed immunogenicity. This challenge drives the continued efforts to deimmunize these protein therapeutics while maintaining their pharmacokinetic properties and therapeutic efficacy. Because mAbs hold a central position in therapeutic strategies against an array of diseases, the importance of conducting comprehensive immunogenicity risk assessment during the drug development process cannot be overstated. Such assessment necessitates the employment of in silico, in vitro, and in vivo strategies to evaluate the immunogenicity risk of mAbs. Understanding the intricacies of the mechanisms that drive mAb immunogenicity is crucial to improving their therapeutic efficacy and safety and developing the most effective strategies to determine and mitigate their immunogenic risk. This review highlights recent advances in immunogenicity prediction strategies, with a focus on protein engineering strategies used throughout development to reduce immunogenicity.
... (Amann et al., 2018(Amann et al., , 2019Bydlinski et al., 2018;Chan et al., 2016;Chung, Wang, Yang, Ponce, et al., 2017;Chung, Wang, Yang, Yin, et al., 2017;Schulz et al., 2018;Tian et al., 2019;Yuan et al., 2019). In mammalian cells, the 1,6-fucosyltransferase (encoded by the FUT8 gene) is the predominant enzyme that transfers fucose to the innermost GlcNAc residue on an N-glycan through a −1,6 linkage Sun et al., 2015;G. Yang, Wang, et al., 2021). ...
Article
Full-text available
For industrial production of recombinant protein biopharmaceuticals, Chinese hamster ovary (CHO) cells represent the most widely adopted host cell system, owing to their capacity to produce high‐quality biologics with human‐like posttranslational modifications. As opposed to random integration, targeted genome editing in genomic safe harbor sites has offered CHO cell line engineering a new perspective, ensuring production consistency in long‐term culture and high biotherapeutic expression levels. Corresponding the remarkable advancements in knowledge of CRISPR‐Cas systems, the use of CRISPR‐Cas technology along with the donor design strategies has been pushed into increasing novel scenarios in cell line engineering, allowing scientists to modify mammalian genomes such as CHO cell line quickly, readily, and efficiently. Depending on the strategies and production requirements, the gene of interest can also be incorporated at single or multiple loci. This review will give a gist of all the most fundamental recent advancements in CHO cell line development, such as different cell line engineering approaches along with donor design strategies for targeted integration of the desired construct into genomic hot spots, which could ultimately lead to the fast‐track product development process with consistent, improved product yield and quality.
... Afucosylation presents one of the most popular glycoengineering approaches for biologics (Lalonde and Durocher, 2017). Afucosylation has been achieved by knocking out FUT8 or the other enzymes involved in the biosynthesis of guanine diphosphate-fucose (Cristea et al., 2013;Sun et al., 2015). ...
Article
Full-text available
The α1,6-fucosyltransferase (encoded by FUT8 gene) is the key enzyme transferring fucose to the innermost GlcNAc residue on an N-glycan through an α-1,6 linkage in the mammalian cells. The presence of core fucose on antibody Fc region can inhibit antibody-dependent cellular cytotoxicity (ADCC) and reduce antibody therapeutic efficiency in vivo. Chinese hamster ovary (CHO) cells are the predominant production platform in biopharmaceutical manufacturing. Therefore, the generation of FUT8 knock-out (FUT8KO) CHO cell line is favorable and can be applied to produce completely non-fucosylated antibodies. The characterization of monoclonal antibodies as well as host cell glycoprotein impurities are required for quality control purposes under regulation rules. To understand the role of FUT8 in the glycosylation of CHO cells, we generated a FUT8 knock-out CHO cell line and performed a large-scale glycoproteomics to characterize the FUT8KO and wild-type (WT) CHO cells. The glycopeptides were enriched by hydrophilic chromatography and fractionated 25 fractions by bRPLC followed by analysis using high-resolution liquid chromatography mass spectrometry (LC-MS). A total of 7,127 unique N-linked glycosite-containing intact glycopeptides (IGPs), 928 glycosites, and 442 glycoproteins were identified from FUT8KO and WT CHO cells. Moreover, 28.62% in 442 identified glycoproteins and 26.69% in 928 identified glycosites were significantly changed in the FUT8KO CHO compared to wild-type CHO cells. The relative abundance of all the three N-glycan types (high-mannose, hybrid, and complex) was determined in FUT8KO comparing to wild-type CHO cells. Furthermore, a decrease in fucosylation content was observed in FUT8KO cells, in which core-fucosylated glycans almost disappeared as an effect of FUT8 gene knockout. Meantime, a total of 51 glycosylation-related enzymes were also quantified in these two cell types and 16 of them were significantly altered in the FUT8KO cells, in which sialyltransferases and glucosyltransferases were sharply decreased. These glycoproteomic results revealed that the knock-out of FUT8 not only influenced the core-fucosylation of proteins but also altered other glycosylation synthesis processes and changed the relative abundance of protein glycosylation.
Article
Chinese hamster ovary (CHO) cells are among the most widely used mammalian cell lines in the biopharmaceutical industry. Therefore, it is not surprising that significant efforts have been made around the engineering of CHO cells using genetic engineering methods such as the CRISPR-Cas system. In this review, we summarize key recent studies that have used different CRISPR-Cas systems such as Cas9, Cas13 or dCas9 fused with effector domains to improve recombinant protein (r-protein) production in CHO cells. Here, every relevant stage of production was considered, underscoring the advantages and limitations of these systems, as well as discussing their bottlenecks and probable solutions. A special emphasis was given on how these systems could disrupt and/or regulate genes related to glycan composition, which has relevant effects over r-protein properties and in vivo activity. Furthermore, the related promising future applications of CRISPR to achieve a tunable, reversible, or highly stable editing of CHO cells are discussed. Overall, the studies covered in this review show that despite the complexity of mammalian cells, the synthetic biology community has developed many mature strategies to improve r-protein production using CHO cells. In this regard, CRISPR-Cas technology clearly provides efficient and flexible genetic manipulation and allows for the generation of more productive CHO cell lines, leading to more cost-efficient production of biopharmaceuticals, however, there is still a need for many emerging techniques in CRISPR to be reported in CHO cells; therefore, more research in these cells is needed to realize the full potential of this technology.
Article
The high demand for monoclonal antibody (mAb) therapeutics in recent years has resulted in significant efforts to improve their costly manufacturing process. The high cost of manufacturing mAbs derives mainly from the purification process, which contributes to 50–80% of the total manufacturing cost. One of the main challenges facing industry at the purification stage is the clearance of host cell proteins (HCPs) that are produced and often co‐purified with the desired mAb product. One of the issues HCPs can cause is the degradation of the final mAb protein product. In this review, we consider techniques that can be used at different stages (upstream and downstream) of mAb manufacture to improve HCP clearance. In addition to established techniques, we also discuss the many new approaches for HCP removal that have the potential to replace current methods for improving HCP reduction and thereby the quality and stability of the final mAb product. This article is protected by copyright. All rights reserved
Chapter
Therapeutic monoclonal antibodies (mAbs) are glycoproteins that contain a pair of N-glycosylation sites on their constant fragment and are widely prescribed for the treatment of several types of cancer and autoimmune disorders. Glycosylation is considered a critical quality attribute of mAbs because it is essential to the safety, pharmacokinetics, and pharmacodynamics of these life-saving biopharmaceuticals. High degrees of glycosylation variability have been observed across different production campaigns of the same mAb product and arise from the numerous biological reactions involved in the glycosylation process, their sensitivity to cell culture conditions, and the genetic background of the production host. Due to its influence in defining the quality of mAb products, substantial effort has been made to develop strategies that minimise mAb glycosylation heterogeneity. This chapter recapitulates the progress made towards controlling mAb glycosylation within the context of biopharmaceutical quality assurance. The chapter presents a critical review of the vast number of (i) cellular, (ii) metabolic, and (iii) in vitro glycoengineering strategies that have been developed to enhance the quality of therapeutic mAbs. We conclude by outlining how these strategies can be combined to achieve the complex task of manufacturing homogenous mAb glycoforms that elicit optimal therapeutic outcomes in the clinic.KeywordsN-linked glycosylationGlycoengineeringTherapeutics antibodiesCell EngineeringMetabolic glycoengineering in vitro glycoengineering
Chapter
The impact of the glycan distribution on the in vivo function and half-life of monoclonal antibodiesMonoclonal antibodies has long motivated the genetic engineering of producer cells to achieve structures that enhance efficacy, safety and stability. To facilitate glycoengineering ofGlycoengineering IgG-producing Chinese hamster ovary cells, we present a rapid protocol that involves the use of RNA interferenceRNA interference (RNAi) for the knockdown of genes of interest coupled with capillary gel electrophoresisCapillary gel electrophoresis and laser-induced fluorescence detection (CGE-LIF) for fast, high-throughput glycan analysis. We apply this methodology to the Fut8 gene, responsible for the addition of core fucose, which is a typical target for increasing antibody-dependent cellular cytotoxicity.
Article
With the emergence of new CRISPR/dCas9 tools that enable site specific modulation of DNA methylation and histone modifications, more detailed investigations of the contribution of epigenetic regulation to the precise phenotype of cells in culture, including recombinant production subclones, is now possible. These also allow a wide range of applications in metabolic engineering once the impact of such epigenetic modifications on the chromatin state is available. In this study, enhanced DNA methylation tools were targeted to a recombinant viral promoter (CMV), an endogenous promoter that is silenced in its native state in CHO cells, but had been reactivated previously (β-galactoside α-2,6-sialyltransferase 1) and an active endogenous promoter (α-1,6-fucosyltransferase), respectively. Comparative ChIP-analysis of histone modifications revealed a general loss of active promoter histone marks and the acquisition of distinct repressive heterochromatin marks after targeted methylation. On the other hand, targeted demethylation resulted in autologous acquisition of active promoter histone marks and loss of repressive heterochromatin marks. These data suggest that DNA methylation directs the removal or deposition of specific histone marks associated with either active, poised or silenced chromatin. Moreover, we show that de novo methylation of the CMV promoter results in reduced transgene expression in CHO cells. Although targeted DNA methylation is not efficient, the transgene is repressed, thus offering an explanation for seemingly conflicting reports about the source of CMV promoter instability in CHO cells. Importantly, modulation of epigenetic marks enables to nudge the cell into a specific gene expression pattern or phenotype, which is stabilized in the cell by autologous addition of further epigenetic marks. Such engineering strategies have the added advantage of being reversible and potentially tunable to not only turn on or off a targeted gene, but also to achieve the setting of a desirable expression level.
Article
Full-text available
Oligosaccharides, sequences of carbohydrates conjugated to proteins and lipids, are arguably the most abundant and structurally diverse class of molecules. Fucosylation is one of the most important oligosaccharide modifications involved in cancer and inflammation. Recent advances in glycomics have identified several types of glyco-biomarkers containing fucosylation that are linked to certain types of cancer. Fucosylated alpha-fetoprotein (AFP) is widely used in the diagnosis of hepatocellular carcinoma because it is more specific than alpha-fetoprotein. High levels of fucosylated haptoglobin have also been found in sera of patients with various carcinomas. We have recently established a simple lectin-antibody ELISA to measure fucosylated haptoglobin and to investigate its clinical use. Cellular fucosylation is dependent upon fucosyltransferase activity and the level of its donor substrate, guanosine diphosphate (GDP)-fucose. GDP-mannose-4,6-dehydratase (GMDS) is a key enzyme involved in the synthesis of GDP-fucose. Mutations of GMDS found in colon cancer cells induced a malignant phenotype, leading to rapid growth in athymic mice resistant to natural killer cells. This review describes the role of fucosylated haptoglobin as a cancer biomarker, and discusses the possible biological role of fucosylation in cancer development.
Article
Full-text available
Chinese hamster ovary (CHO) cells are widely used in the biopharmaceutical industry as a host for the production of complex pharmaceutical proteins. Thus genome engineering of CHO cells for improved product quality and yield is of great interest. Here, we demonstrate for the first time the efficacy of the CRISPR Cas9 technology in CHO cells by generating site-specific gene disruptions in COSMC and FUT8, which encode proteins involved in glycosylation. The tested single guide RNAs (sgRNAs) created an indel frequency up to 47.3% in COSMC, while an indel frequency up to 99.7% in FUT8 was achieved by applying lectin selection. All eight sgRNAs examined in this study resulted in relatively high indel frequencies, demonstrating that the Cas9 system is a robust and efficient genome-editing methodology in CHO cells. Deep sequencing revealed that 85% of the indels created by Cas9 resulted in frameshift mutations at the target sites, with a strong preference for single base indels. Finally, we have developed a user-friendly bioinformatics tool, named “CRISPy” for rapid identification of sgRNA target sequences in the CHO-K1 genome. The CRISPy tool identified 1,970,449 CRISPR targets divided into 27,553 genes and lists the number of off-target sites in the genome. In conclusion, the proven functionality of Cas9 to edit CHO genomes combined with our CRISPy database has the potential to accelerate genome editing and synthetic biology efforts in CHO cells. Biotechnol. Bioeng. © 2014 The Authors. Biotechnology and Bioengineering Published by Wiley Periodicals, Inc.
Article
Full-text available
Monomeric CRISPR-Cas9 nucleases are widely used for targeted genome editing but can induce unwanted off-target mutations with high frequencies. Here we describe dimeric RNA-guided FokI nucleases (RFNs) that can recognize extended sequences and edit endogenous genes with high efficiencies in human cells. RFN cleavage activity depends strictly on the binding of two guide RNAs (gRNAs) to DNA with a defined spacing and orientation substantially reducing the likelihood that a suitable target site will occur more than once in the genome and therefore improving specificities relative to wild-type Cas9 monomers. RFNs guided by a single gRNA generally induce lower levels of unwanted mutations than matched monomeric Cas9 nickases. In addition, we describe a simple method for expressing multiple gRNAs bearing any 5' end nucleotide, which gives dimeric RFNs a broad targeting range. RFNs combine the ease of RNA-based targeting with the specificity enhancement inherent to dimerization and are likely to be useful in applications that require highly precise genome editing.
Article
Full-text available
Since the discovery of hybridoma cells by Kohler and Milstein, the uses of monoclonal antibody (mAb), the protein produced by such cells are in vogue. Such antibodies with single isotype have higher specificity, and the serological tests employied in show higher reproducibility compared to those with use of poly-clonal antisera. There are several procedures of mAb production which vary considerably but the principle remains the same which states that antigens intro-duced into animals generally result in the stimulation of lymphocytes, some of which produce antibody of only one type, although the isotype may change. The developments in the field of cell culture and transfec-tion technology have lead to the production of im-proved qualities of mAbs. In general, monoclonal an-tibodies are important reagents used in biomedical research, such as, in the field of diagnostics and therapeutics as well as targeted drug delivery systems. They have got importance not only for infectious dis-eases caused by microbes and parasites, but also for cancer, metabolic and hormonal disorders, in the di-agnosis of lymphoid and myeloid malignancies and tissue typing, enzyme linked immunosorbent assay (ELISA) (especially blocking ELISA), radio immu-noassay (RIA), serotyping of pathogens and their immunological intervention with passive antibody, anti-idiotype inhibition or magic bullet therapy with cytotoxic agents coupled with antimouse specific an- tibody. The application of mAbs in diagnosis of vari-ous livestock diseases is an important area of concern as these diseases are a major and increasingly impor- tant factor reducing livestock productivity in various parts of the world. In this context, the application of mAbs for diagnosis of important bacterial diseases viz., Anthrax, Brucellosis, Paratuberculosis, Lepto-spirosis, Listeriosis, Clostridial infections and My-coplasmosis (CBPP), fungal diseases viz., Zygomyco-sis, Cryptococcosis, Histoplasmosis and Paracoccidi-odomycosis, viral diseases viz., Foot-and-mouth dis-ease (FMD), Infectious bovine rhinotracheitis/Infec- tious pustular vulvovaginitis (IBR/IPV), Rota viral diarrhoea, Blue tongue, Rabies, Classical swine fever and re-emerging viral diseases like Hendra and Nipah viral infections and parasitic diseases viz., dirofilariosis, and Trichinellosis and haemportozoan diseases (in-cluding Trypanosomiasis, Leishmaniasis, Anaplas-mosis, infections caused by Plasmodium spp. as well as tick borne diseases) have been discussed thor-oughly along with the specifications of the diagnostic assays for each disease for the convenience of the di-agnosticians, researchers, scientists and students to employ such assays, both in field and laboratories to strengthen the disease control programme.
Article
Recent advances in genome engineering technologies based on the CRISPR-associated RNA-guided endonuclease Cas9 are enabling the systematic interrogation of mammalian genome function. Analogous to the search function in modern word processors, Cas9 can be guided to specific locations within complex genomes by a short RNA search string. Using this system, DNA sequences within the endogenous genome and their functional outputs are now easily edited or modulated in virtually any organism of choice. Cas9-mediated genetic perturbation is simple and scalable, empowering researchers to elucidate the functional organization of the genome at the systems level and establish causal linkages between genetic variations and biological phenotypes. In this Review, we describe the development and applications of Cas9 for a variety of research or translational applications while highlighting challenges as well as future directions. Derived from a remarkable microbial defense system, Cas9 is driving innovative applications from basic biology to biotechnology and medicine.
Article
Targeted nucleases are powerful tools for mediating genome alteration with high precision. The RNA-guided Cas9 nuclease from the microbial clustered regularly interspaced short palindromic repeats (CRISPR) adaptive immune system can be used to facilitate efficient genome engineering in eukaryotic cells by simply specifying a 20-nt targeting sequence within its guide RNA. Here we describe a set of tools for Cas9-mediated genome editing via nonhomologous end joining (NHEJ) or homology-directed repair (HDR) in mammalian cells, as well as generation of modified cell lines for downstream functional studies. To minimize off-target cleavage, we further describe a double-nicking strategy using the Cas9 nickase mutant with paired guide RNAs. This protocol provides experimentally derived guidelines for the selection of target sites, evaluation of cleavage efficiency and analysis of off-target activity. Beginning with target design, gene modifications can be achieved within as little as 1-2 weeks, and modified clonal cell lines can be derived within 2-3 weeks.
Article
Targeted genome engineering has been instrumental for the study of biological processes, and it holds great promise for the treatment of disease. Historically, gene targeting by homologous recombination has been the preferred method to modify specific genes in mouse and human cells (Fig. 1). However, this approach is hampered by low efficiency, the requirement for drug selection to detect targeted cells, and the limited number of cell types and organisms amenable to the method. To overcome these limitations, technologies based on sequence-specific zinc finger (ZF) and transcription activator-like effector (TALE) proteins have been developed. These proteins can be engineered to theoretically recognize any DNA sequence in the genome. When fused to a nuclease domain and assembled in pairs flanking a target site of interest, ZF and TALE nucleases (ZFNs and TALENs) will introduce double-strand breaks (DSBs) on DNA. DSBs serve as substrates for nonhomologous end joining (NHEJ) or homology-directed repair (HDR), which in turn facilitate the engineering of targeted mutations, repair of endogenous mutations, or introduction of transgenic DNA elements. The clustered, regularly interspaced, short palindromic repeat (CRISPR)-CRISPR-associated (Cas) system represents the latest addition to this arsenal of tools for site-specific genome engineering (see below). Although each of these three gene modification approaches has advantages and disadvantages (Fig. 1), the pace and ease with which the CRISPR-Cas systems have been adapted to modify genes in different cell types and organisms suggests that it may very well become the new method of choice for genome engineering. In PNAS, Hou et al. introduce a unique variant of the Cas9 enzyme (1), which provides additional flexibility and specificity to the CRISPR system and to genome-modifying tools in general.
Article
Targeted genome editing technologies have enabled a broad range of research and medical applications. The Cas9 nuclease from the microbial CRISPR-Cas system is targeted to specific genomic loci by a 20 nt guide sequence, which can tolerate certain mismatches to the DNA target and thereby promote undesired off-target mutagenesis. Here, we describe an approach that combines a Cas9 nickase mutant with paired guide RNAs to introduce targeted double-strand breaks. Because individual nicks in the genome are repaired with high fidelity, simultaneous nicking via appropriately offset guide RNAs is required for double-stranded breaks and extends the number of specifically recognized bases for target cleavage. We demonstrate that using paired nicking can reduce off-target activity by 50- to 1,500-fold in cell lines and to facilitate gene knockout in mouse zygotes without sacrificing on-target cleavage efficiency. This versatile strategy enables a wide variety of genome editing applications that require high specificity.
Article
Clustered, regularly interspaced, short palindromic repeat (CRISPR) RNA-guided nucleases (RGNs) have rapidly emerged as a facile and efficient platform for genome editing. Here, we use a human cell-based reporter assay to characterize off-target cleavage of CRISPR-associated (Cas)9-based RGNs. We find that single and double mismatches are tolerated to varying degrees depending on their position along the guide RNA (gRNA)-DNA interface. We also readily detected off-target alterations induced by four out of six RGNs targeted to endogenous loci in human cells by examination of partially mismatched sites. The off-target sites we identified harbored up to five mismatches and many were mutagenized with frequencies comparable to (or higher than) those observed at the intended on-target site. Our work demonstrates that RGNs can be highly active even with imperfectly matched RNA-DNA interfaces in human cells, a finding that might confound their use in research and therapeutic applications.
Article
The carbohydrate composition of lignocellulosic biomass hydrolysates is highly complex. High performance anion exchange chromatography coupled with pulsed amperometric detection (HPAEC-PAD), a widely used method for carbohydrate analysis, provides limited chemical information on the detected peaks. To improve the detection and increase the chemical information of the carbohydrates, HPAEC was coupled with mass spectrometry (MS). Using a pooled hydrolysate sample, it was shown that HPAEC-MS can separate and detect many oligosaccharides in one experimental run based on retention time and mass. The method was validated on its linearity, reproducibility and response factors. The analysis of a group of different biomass hydrolysates revealed that remaining disaccharides was the bottleneck of the hydrolysis process. As an analytical tool, HPAEC-MS provides information for the improvement of hydrolysate pretreatment method and enzyme cocktail quality. Besides, the consumption ability of microbial host strains for various mono- and oligosaccharides in hydrolysates can be assessed.