103

Optimal Human Embryo Culture Jason E. Swain, PhD, HCLD1,2 1 National Foundation for Fertility Research, Colorado Center for

Reproductive Medicine, Lone Tree, Colorado 2 Fertility Lab Sciences, Englewood, Colorado

Address for correspondence Jason E. Swain, PhD, HCLD, 10290 RidgeGate Circle, Lone Tree, CO 80124 (e-mail: [email protected]).

Abstract

Keywords

► ► ► ► ►

optimal culture embryo blastocyst IVF

A large contributor to success during in vitro fertilization (IVF) lies in the processes occurring within the IVF laboratory. These processes make up the “culture system.” This system entails numerous procedures and technical steps that must be optimized to produce a competent embryo. Notably, variations exist between programs that include differences in patient population, clinical stimulation, and other factors. Thus, a single “optimal” culture system to be utilized between all laboratories is likely not feasible. Rather, laboratory procedures should be optimized based on an individual laboratory’s performance. That being said, within the scientific literature, there are key components, approaches, and techniques within the culture system that have been shown to be superior to alternatives. These key components important in improving embryo culture are discussed.

As the number of IVF cycles continues to increase worldwide, there is an obvious desire to continue to improve upon the rising success rates, as well as system efficiency. This not only improves the patient’s chance at achieving a successful pregnancy in a timely fashion, but also presumably makes the process more readily available to the wider population. There are multiple components in the IVF process that impact success. These include, but are not limited to, optimizing oocyte quality through appropriate stimulation, sperm quality, and ensuring appropriate uterine receptivity. However, despite these clinical aspects or the inherent properties of patients themselves, immense focus and responsibility is often placed on the laboratory component of the IVF process. It is often viewed that much of what drives IVF success occurs within the confines of the IVF laboratory. Certainly, when embryo development is poor or outcomes are low, one of the first suspected culprits is a suboptimal culture system. Indeed, laboratory conditions are one component of the system than can impact embryo quality. As a result, immense amounts of time and resources are devoted to developing the “optimal” system to culture embryos. The key word here, however, is “system,” as culturing embryos consists of several variables. Refining or optimizing each of these components is required to maximize embryo development and resulting outcomes. Furthermore, many products

Issue Theme Best Practices in In Vitro Fertilization; Guest Editor, Bradley J. Van Voorhis, MD

and approaches are developed using animal models and determining whether a product is “optimized” for human embryos can be difficult. Importantly, within the context of the entire culture system, “optimal” conditions may vary slightly from laboratory to laboratory. However, initial gamete quality aside and accounting for laboratory-specific considerations, there are certain critical components of the culture system that are important to maximizing resulting embryo development.

Gamete Processing As mentioned, embryo quality begins with the gametes. A significant component of gamete quality is inherent and beyond influence of laboratory conditions. Thus, it should be mentioned, that if gamete quality is poor, even an “optimal” culture system is likely unable to overcome inherent abnormalities. That being said, sperm and oocytes are cultured briefly within the laboratory and these brief periods can impact quality. Final maturation of oocytes in vitro prior to insemination/ injection is critical to maximize embryo development. Oocytes are purposely retrieved prior to ovulation at approximately 36 hours post–human chorionic gonadotropin (hCG), and though many may appear mature via presence of the first

Copyright © 2015 by Thieme Medical Publishers, Inc., 333 Seventh Avenue, New York, NY 10001, USA. Tel: +1(212) 584-4662.

DOI http://dx.doi.org/ 10.1055/s-0035-1546423. ISSN 1526-8004.

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

Semin Reprod Med 2015;33:103–117

Optimal Human Embryo Culture

Swain

polar body, nuclear and cytoplasmic maturation are likely still incomplete. Polarized microscopy has demonstrated that in vitro matured oocytes displaying extrusion of the first polar body may still be immature; and actually be at the telophase stage of meiosis. True metaphase II spindle formation, or completion of nuclear maturation, does not occur until approximately 1 hour following polar body extrusion.1 In these in vitro matured eggs, injection less than 1 hour after polar body extrusion yielded significantly lower fertilization rates compared with other timings, with the highest success coming with injection occurring 2 to 4 hours after retrieval.1 This suggests that a brief “final maturation” may be beneficial prior to intracytoplasmic sperm injection (ICSI). Supporting this approach, when examining different injection timings following a 36-hour post-hCG retrieval, it was found that culture of cumulus oocyte complexes (COCs) for 3 to 6 hours prior to injection yielded superior fertilization and cleavage embryo development compared with COCs cultured less than 3 hours.2 Similarly, using patient randomization in a prospective trial, waiting 4 hours to inject oocytes following retrieval yielded superior fertilization and blastocyst development compared with injection immediately after retrieval.3 Thus, maturation of oocytes for 3 to 4 hours following retrieval is often recommended. Though minimal trimming of cumulus cells to remove blood clots is likely warranted at time of retrieval to avoid extended culture in their presence and possible detrimental effects of reactive oxygen species (ROS) or interference with sperm during standard insemination, it does not appear to improve oocyte quality or improve ICSI outcomes4,5 and final maturation of oocytes should occur with cumulus cells largely intact. Maturing oocytes for the final 4 hours following retrieval with cumulus cells intact appeared to yield superior fertilization and blastocyst development compared to those that were denuded immediately after retrieval3 (►Table 1). It should be noted that at least one other study has shown no difference in fertilization or cleavage embryo development following ICSI with when comparing final maturation timings of 1 to 2 hours versus 5 to 6 hours or with cumulus intact versus denuded oocytes.6 It is unknown if waiting any period of time following denuding oocytes before injection is beneficial. Final oocyte maturation following retrieval commonly occurs in a specialized fertilization media with elevated glucose concentrations, commonly formulated based on met-

abolic needs of sperm and/or cumulus-oocyte-complexes. This may be more important for cases involving standard insemination than ICSI. Alternatively, final oocyte maturation can also be performed within embryo culture media. It is unclear whether a specific media system for final oocyte maturation is superior, and differences in media formulation may be one partial explanation for contradictions in the literature regarding the impact of final oocyte maturation and specific time windows. When trying to maintain gamete quality to ensure optimal embryo development within the culture system, care should also be taken when denuding oocytes. It has been shown that mechanical manipulations during denuding can displace the polar body,7,8 thereby making it difficult to determine the location of the meiotic spindle, which can be problematic in cases of ICSI. Therefore, limiting the size of the pipet used to denude oocytes is prudent. Sizes of approximately 140 µm are sufficiently small to remove the majority of cumulus to permit ICSI and other micromanipulation procedures while avoiding undue physical stress to the egg. Additionally, at least in terms of denuding oocytes in cases of ICSI, limiting exposure to hyaluronidase is likely warranted. Common recommendations include 45 to 60 seconds, taking care to wash oocytes immediately and thoroughly afterward. At least one peer reviewed study using a retrospective analysis without sibling oocytes splits indicates that a recombinant hyaluronidase is superior to bovine-derived enzyme, yielding higher fertilization rates and lower oocyte damage.9 A prospective, randomized trial using sibling oocytes splits also showed a recombinant hyaluronidase yielded superior fertilization compared with bovine hyaluronidase,10 while another trial using sibling oocytes splits indicated noninferiority to animal-derived enzyme.11 Detailed discussion regarding sperm preparation and impact on embryo quality is beyond the scope of this review. It is known that separation and processing techniques can impact sperm quality and DNA integrity and likely resulting embryo quality.12,13 Various studies exist comparing common practices such as density gradient separation and swim-up and impact on sperm. A clear consensus as to the superior approach is not clear from the literature, and conflicting results may stem from subtle variations in techniques, such as gradient used, centrifugation speed and time of centrifugation, number of washes, media used, swim-up technique,

Table 1 Impact of final oocyte maturation timing and denuding prior to ICSI and impact on outcomes Endpoint

Immediate Denuding & Injection

Delayed Denuding & Injection

Immediate Denuding & Delayed Injection

# MII Oocytes

319

313

315

a

72.5

b

46a,b

% Fertilization

43.4

% Blastocysts

16.7aþ

42.9bþ

18.8a,b

% Clinical pregnancy

8.7

26.2

11.8

Abbreviation: ICSI, intracytoplasmic sperm injection. Note: Different letter superscripts represent differences between treatment groups.  p < 0.05. þ p < 0.06.

Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

104

Swain

105

Abbreviation, CPR, clinical pregnancy rate. a Paternot et al (2010) used different handling media, fertilization media between groups.

ISM1 vs. GM501 Campo et al (2010)

83 vs. 84

Day 2/3

26.5 vs. 22.6%

17.8 vs. 20.0

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

17.8 vs. 20.0

36.8 vs. 57.5 (p < 0.08) Day 5/6 ECM/Multiblast vs. Global Supelveda et al (2009)

38 vs. 40

Sydney IVF cleavage vs. GM501 Paternot et al (2010)

70 vs. 77

Day 3 or 5

52.6 vs. 72.5

36.8 vs. 57.5 (p < 0.08)

19.6 vs. 26%

CPR/transfer (seq vs. single) Transfer day (2/3 vs. 5/6) Total transfers Media compared Study

Table 2 Comparisons of sequential and single-step media using patient randomization

In trying to discern key components of the “optimal” embryo culture system, focus immediately turns to culture media. Unquestionably, culture media plays a significant role in embryo development and quality, and therefore clinical outcomes. However, in terms of identifying or selecting the single “optimal” medium for use within IVF laboratories, this is likely a futile endeavor. Numerous variables present within the laboratory, many discussed later in this article, impact culture media performance.21 Thus, the same medium can perform differently between laboratories. The inability to identity a single optimum media is reflected by examining the numerous studies in the literature comparing one medium against another.22 Indeed, when examining prospective trials using sibling oocytes splits or patient randomization, comparison of different sequential media23–29 and comparisons between single-step culture media and sequential media30–34 fail to identify a superior approach or product (►Tables 2–5). These studies are often underpowered and not rigorously controlled, thus making it difficult to draw definitive conclusions. Indeed, no trial exists comparing every available media against each other in a controlled fashion, which would be required to identify the superior medium. It is also unclear whether changing media at 48- or 72-hour intervals, or utilizing uninterrupted culture yields superior results with modern media formulations. Individual laboratories should perform their own trials/splits to determine which product and methodology performs most successfully under their specific conditions. That being said, when selecting a medium, laboratories can make sure to consider key factors. Carbohydrates: Basal culture media will contain pyruvate, lactate, and glucose in various combinations and concentrations. This will vary from medium to medium and no clear consensus on an optimal concentration or ratio of components exists. Pyruvate can be labile in solution35,36 and may be influenced by alkaline pH, which has led some to adopt a practice of limiting the time frame they will use a media once a bottle has been opened. A proper ratio of lactate to pyruvate is likely important to maintain cellular redox potential and is often recommended that media only contain the metabolizable L-form of lactate, rather than mixture of L- and D-forms, to avoid unnecessary lowering of cellular internal pH.37

Implantation rate (total)

Culture Media

20 vs. 28%

Implantation rate day 3 (seq vs. single)

Implantation rate day 5/6 (seq vs. single)

and other variables.14 Appropriate sperm concentrations should be used for standard insemination cases, with 50,000 to 100,000 motile sperm/mL as standard accepted concentrations. Elevated sperm concentrations could be detrimental due to increased ROS, lowered pH, or other media alterations. In regard to ICSI, at a minimum, normal morphology sperm should be utilized for injection with sperm assessment at 400x. Use of higher magnification approaches, such as intracytoplasmic morphological sperm injection (IMSI) do not appear overly beneficial, though this may depend on patient population.15–18 Emerging selection approaches, such as those use hyaluronan binding to identify/select sperm, may also be useful,19 though this likely depends on the patient population and semen characteristics.20

20 vs. 43%a

Optimal Human Embryo Culture

Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

Seminars in Reproductive Medicine

Vol. 33

17 vs. 25 (31 had embryos from both)

13 vs. 17 (19 had embryos from both)

7 vs. 21 (52 had embryos from both)

No. of transfers for each media

Day 3

Day 3–5

Day 3 or 5

Transfer day (3 vs. 4/5)

58.8 vs. 48%

0 vs. 40% (0/1 vs. 2/5)

CPR D3 (seq vs. single) a

66.6 vs. 58.3%

CPR day 4/5 (seq vs. single)a

38.2 vs. 42%

0 vs. 28.6%

Implantation rate D3 (seq vs. single)

No. 2/2015

P1 þ 20% SSS vs. IVF-50 þ 10 mg/mL HSA

P1 þ 20% SSS vs. Cook IVF þ 1% HSA

G1.1/G1.2 vs. Sydney IVF Cleavage/Blastocyst

G1.2/G2.2 vs. GIII series

G1.2/G2.2 vs. BlastAssist M1/M2

Mauri et al (2001)a

Ben-Yosef et al (2004)b

Van Langendonckt et al (2001)

Balaban and Urman (2005)c

Zollner et al (2004)d

182 (196 cycle) vs. 167 (179 cycles)

91 each media

Patients/cycles

Day 3–5

Day 3 or 5

Day 3 or 5/6

Day2/3

Day 3

Transfer day (2/3 vs. 4/5/6)

Overall CPR: 31.0 vs. 28.1%

37.8 vs. 50.3 (p < 0.01)

48.6 vs. 48.7

19.4 vs. 13.8

37.0 vs. 33.7

CPR/transfer day 3

51.4 vs. 69.4

48.6 vs. 48.7

CPR/transfer day 5/6

Overall implantation: 13.2 vs. 13.0%

14.5 vs. 25.7 (p < 0.05)

18.9 vs. 16.6

9.9 vs. 6.0

17.4 vs. 15.8

Implantation rate day 3 transfer

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

Abbreviation, CPR, clinical pregnancy rate. a Mauri et al (2001) also changed sperm prep media, PVP, HEPES-media, and oil for each of the treatment groups. b Ben-Yosef et al (2004) also changed sperm preparation and oocyte incubation media between treatment groups. c Balaban and Urman (2005) changed handling media and other preparatory steps between treatment groups. d Zollner et al (2004) did not report CPR or implantation based on day of transfer, but did note that no statistically significant differences existed between day of transfer or media used.

Media compared

Study

Table 4 Comparison of different sequential media systems using patient randomization

29.0 vs. 45.1 (p < 0.05)

43.1 vs. 36.1

Implantation rate day 5/6 transfer

57.1 vs. 52.6%

Implantation date D4/5 (seq vs. single)

Clinical pregnancy rate (CPR) and implantation rates were calculated using only cycles in which all embryos were transferred from one media group. Single-step media of Reed et al did not have any media refreshment during day 5 culture. The study did not permit reporting of CPR or implantation rates for single versus sequential media.

73

49

80

No. of patients (total transfers)

Optimal Human Embryo Culture

b

a

ECM/multiblast vs. SSM

Ciray et al (2012)

Sage vs. Global

G5 Series vs. Global

Reed et al (2009)b

Basile et al (2013)

Media compared

Study

Table 3 Comparison of sequential and single-step media using sibling embryo split

106 Swain

Optimal Human Embryo Culture

Swain

107

Table 5 Comparison of sequential media systems using sibling embryo splits Study

Media compared

No. of patients (total transfers)

No. of transfers for each media

Transfer day (2/3 vs. 5/6)

CPR/transfera

Implantation rate (seq vs. single)

Hambiliki et al (2011)b

G1.5 vs. EmbryoAssist

108

69 vs. 22 (17 had embryos from both)

Day 2/3

46.4 vs. 36.4

40.9 vs. 37.5

Sifer et al (2009)

GIII vs. ISM1

190 vs. 102 (298 had embryos from both)

Day 3

44.2 vs. 35.3

32.2 vs. 28.4

Glucose can be present or absent for early cleavage development38,39 but is required for later stages to promote compaction and blastocyst formation. Amino acids: Amino acids support numerous cellular processes, including acting as metabolites, osmolytes, antioxidants, and buffers, and likely help alleviate stress in the embryo culture system.40 Brief period of culture with no amino acids present can impair mouse embryo development.41 Thus, all embryo culture media should contain some assortment of amino acids. While animal studies have given some insight,42–44 it is likely impossible to determine the optimal mixture and concentrations of specific amino acids for human embryo culture. Studies have identified glycine, taurine, and glutamine as important amino acids for human embryos.45–47 If glutamine is present, modern embryo culture media should include the dipeptide form of the amino acid. Use of the stable dipeptide-glutamine helps avoid ammonium accumulation and the associated detrimental effects on human embryo development, and appears superior to glutamine.48–51 No clear consensus exists regarding whether glycyl-glutamine or alanyl-glutamine is the superior dipeptide for human embryo culture. Antibiotic: Though not required, an antibiotic of some sort should likely be included in embryo culture media. If antibiotic is included, gentamicin, rather than penicillin-streptomycin (pen-strep), should be used. Pen-strep–resistant cases of contamination in IVF52 were 91%, and pen-strep was detrimental to human and hamster embryo development,53,54 possibly via damage to embryo chromatin, gene expression, and increased apoptosis.55 Additionally, gentamicin is heat and pH stable, and of 70 bacterial strains in contaminated embryo media, all were sensitive to gentamicin.52 Protein: Protein supplementation to culture media can have a dramatic impact of embryo development and resulting outcomes. The macromolecule supplementation can act as a surfactant, as a nitrogen source, to stabilize membranes, to modulate physical microenvironment, as a carrier molecule for other compounds, or even to bind trace elements/toxins. A recent report even indicates that protein in culture media can impact human birthweight,56 though results of the study should be interpreted with caution due to lack of control over other system variables. Regardless, it is irrefutable that protein supplements are one of the least defined components of

the culture system, and thus a large source of potential variation. For example, stabilizers/preservatives used when making protein solutions, such as octanoic acid, may be embryotoxic.57 In addition, as mentioned, other proteins, growth factors, and hormone “contaminants” may be present at varying levels in protein preparation.58–60 Some clinical IVF media currently include growth factors, though limited data exist to support their purposeful use61–63 and this remains a controversial topic.64 Thus, to optimize performance of the culture system, specific lots of protein should be screened for efficacy prior to clinical implementation Primary protein supplements used in clinical embryo culture media include human serum albumin (HSA), as well as complex protein supplements containing HSA and a combination of α and β globulins.65,66 Recombinant albumin is also available, as well as a dextran-based macromolecular supplement. Data suggest that a complex protein supplement with globulins is superior to HSA for animal embryos,65,67 human embryo development,68,69 and even live birth.70 This may be due to growth hormones and other components in the complex protein products.59 Recombinant albumin gave comparable embryo development to HSA in a prospective randomized trial and may reduce media variability.71 Optimal concentrations of protein are unknown, though 5% of albumin or 10% of complex protein products are commonly used and higher concentrations of up to 20% of a globulin supplement may be beneficial in certain circumstances.72 Oil: Oil overlay is commonly used during embryo culture to prevent media evaporation, which can result in media osmolality increase and compromised embryo development. Oil is required if using modern benchtop incubators with no humidification. However, open culture, or culture without oil overlay, may also be used if humidification is present and a sufficient volume of media is used. Similar to protein, despite improvement in refinement and testing, oil overlay products can differ significantly due to contaminants and are a large source of variation within the culture system. Washing of oil with either water or culture media can help alleviate some toxicity concerns73 and a good practice to ensure that all oil is washed. Additionally, proper storage is required to avoid peroxidation of oil, which can lead to embryo toxicity.73–76 Recommendations often include keeping oil out of direct light and storing at 4°C. Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

Note: Limitations often exist with media comparisons, including the changing of several variables, making it very difficult to determine impact of the media itself. a Clinical pregnancy rate (CPR) and implantation rates were calculated using only cycles where all embryos were transferred from one media group. b Hambiliki et al (2011) used different fertilization media and different proteins for each group.

Optimal Human Embryo Culture

Swain

There are several oil products for embryo culture, including paraffin oil and light mineral oil. Products are sold in plastic or glass bottles. There is no clear consensus on a superior oil type, and preferences of a particular laboratory may be based more on oil viscosity and impact of dish preparation techniques. Regardless of the oil, testing using an approved bioassay is recommended prior to use. pH: pH of culture media is of special concern, as this parameter is directly controlled by and can vary significantly between laboratories. Media pH is primarily determined by the bicarbonate concentration in the media itself and the CO2 concentration of the culture incubator. However, factors such as laboratory elevation, protein supplementation as well as different media types all impact pH and prevent a single incubator CO2 value from being used throughout the field. Thus, this parameter should be measured and set by each laboratory to account for their particular conditions. Though embryos can develop over a range of media pH, it appears it may impact embryo development. It is known that changes in internal pH can impact embryo metabolism, development, and even fetal growth in rodents.77–79 Certain cell types, such as the denuded mature oocyte or thawed embryos, lack robust internal pH regulation and may be more susceptible to changes in media pH.77,80,81 Preliminary data suggest media pH may impact human embryo development.82,83 However, it is difficult to isolate pH as the sole variable as both bicarbonate or CO2 values change to alter media pH and both can influence embryo biosynthetic pathways. Regardless, maintaining tight control over media pH is likely prudent as part of a rigorous QC program. Unfortunately, no optimal pH has been identified for human embryo culture.84 Phenol red is a common media component added to media to help visually monitor pH changes. This may be useful to some to gauge media age or incubator function, etc. Some media have chosen to exclude this indicator due to possible negative impact on embryo development. It has been suggested that possible estrogenic effects of phenol red could be problematic. However, estrogenic activity appears to be due to the contaminants in certain preparations, not due to the phenol red itself.85 In addition, it has been suggested that phenol red use may lead to ROS generation following light exposure,86 though this has not been shown using modern human embryo culture media or under conditions utilized for modern human embryos. Therefore, it cannot be concluded that phenol red conveys a negative impact, and its use may be warranted in a selected commercial media if a visual pH indicator is useful for a particular laboratory. Related to pH, use of media that include synthetic zwitterionic buffers to stabilize pH for procedures outside the laboratory incubator may impact embryo development. These buffered media are often used for procedures such as oocyte retrieval, sperm preparation, ICSI, embryo transfer, and cryopreservation. Common buffers include HEPES, MOPS, or their combination. Despite poorly designed studies indicating buffers, such as HEPES, may be toxic for human gametes or embryos following ICSI,87 numerous studies have demonstrated safety of both HEPES and MOPS and there Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

is no clear advantage of one over the other in terms of embryo development.88,89 When other variables are controlled, including pH, gas atmosphere, bicarbonate levels, and substrate composition, no detrimental effect of HEPES or MOPS have been observed.88–91 Importantly, phosphate buffered saline (PBS) should not be used for procedures involving gametes and embryos, as concentrations of phosphate required to provide adequate buffering have consistently been shown to impair embryo development.89,91 Of note, if care is taken in the laboratory and proper equipment is used, including gassed isolette incubator chambers, use of these specialized buffered media can be avoided for oocyte retrieval or embryo transfer and may help reduce the number of variables in the culture system. Osmolality: Osmolality is a component of culture media that can significantly impact embryo development. Several studies exist demonstrating that high osmolality can result in developmental arrest of rodent embryos.92–96 The impact of osmolality on embryo development depends on media composition, as components, such as key amino acids, can act as osmolytes and help cells regulate volume. Though osmolality is set by the commercial media manufacturer, conditions in the laboratory can alter osmolality and thereby impact embryo development. Factors impacting evaporation can result in media osmolality shifts and are present during media/dish preparation.97 Time, media volume, airflow, and temperature can all result in detrimental osmolality shifts and compromise embryo development. Thus, a strict regime must be followed when performing daily media preparation that limits these variables to avoid altering media osmolality. Additionally, evaporation within the incubator must also be considered. This is especially important considering modern incubators that are devoid of humidification.

Incubators Minimizing environmentally induced stress within the IVF laboratory is crucial in creating a culture system optimized for embryo development and to achieve maximal assisted reproductive outcomes. Key environmental variables to consider within the culture system include properties of culture media, such as the aforementioned pH and osmolality, as well as variables such temperature and air quality. Importantly, all of these potential stressors can be impacted by the laboratory incubator. As a result, the incubator is arguably the most important piece of equipment in the laboratory. Thus, incubator selection and management is critical to ensuring optimization of the culture system. Multiple incubator types exist with varying capabilities and differing methods of regulating their internal environment. Box-type incubators, initially developed to hold multiple flasks of somatic cells, have been long used for clinical IVF. These were later adapted to smaller box-type incubators. Most recently, a variety of embryo-specific benchtop incubators have been developed. These newer incubators tend to have small chambers and direct heat to try to speed environmental recovery following incubator openings/closings. As a result, selection of an appropriate culture incubator for the

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

108

IVF laboratory has become a complex process. To date, there is no clear consensus as to a superior incubator type, as efficacy and environmental stability rely heavily on incubator use and management.98 Thus, an appropriate number of incubators are required to avoid excessive openings/closings. Reduced incubator opening appears to benefit mouse embryo development,41,99 as well as improve human blastocyst development.100 A mix of incubator types within an individual laboratory permits a wide variety of uses to accommodate different dish or test tube types, as well as emerging culture technology. That being said, there are certain environmental variables that should be utilized within any incubator type. Atmosphere: Today’s optimal culture system requires the use of low oxygen within the incubator. Numerous animal studies; including of the mouse, cat, sheep, pig, cow, and rat, have repeatedly shown the benefit of using reduced oxygen concentrations compared with atmospheric levels. Similarly, use of low oxygen has also been studied extensively in the human and the majority of studies demonstrate an improvement in embryo development.101–112 Admittedly, confounding variables in some experimental designs can sometimes make direct comparisons difficult. That being said, no studies indicate a detriment of using low oxygen. Importantly, use of low oxygen throughout the entire culture period through day 5/6 appears to be required to see the most benefit.107,109,113 In one of the better designed studies, a prospective RCT in a high performing program showed that use of low oxygen significantly improved pregnancy, implantation, and live birth rates (►Table 6).113 Similar improvements in live birth have been reported by others.114 The exact mechanism of the benefit of low oxygen use for embryo culture is unknown. Use of low oxygen may be superior due to improved embryo metabolism, reduced ROS, or perhaps even improved air quality. It is known that air quality within the incubator is another component of the culture system that can impact embryo development.115 The presence of volatile organic compound (VOCs), such as aldehydes or toluene, within the air of IVF laboratories has been well-documented throughout the literature,116–118 and it is well accepted that poor air quality can compromise the culture system, whereas measures to improve air quality result in improved outcomes,119 presumably due to the detrimental effects of this VOCs on the preimplantation embryo.116,118,120,121 Low oxygen incubators use approximately 90% nitrogen or premixed cylinders of gas in place of room air. This may provide purer gas quality within the laboratory incubator as compared to those that use high

Swain

oxygen/room air. However, this may not always be the case, as gas cylinders can be the source of harmful VOCs, such as benzenes or alcohols.116,117 Thus, inclusion of specialized inline filters from the gas supply or filters placed within the incubator can remove VOCs and improve outcomes,121–124 though this is not always the case122,125,126 and may depend on the existing air quality in the laboratory/ incubator. Temperature: Temperature is another variable of the culture system that may impact efficacy. It is well known that temperature can impact various aspects of gamete and embryo function, most notably, meiotic spindle stability127–129 and possibly embryo metabolism.130 Studies have reported that maintaining media temperature around 35 to 37°C compared with 25 to 30°C during retrieval is beneficial for bovine embryo development and mouse embryo gene expression.131,132 Maintaining temperature stability around 37°C during human oocyte injection also improved fertilization rates and embryo development.128 However, the optimal temperature at which to culture embryos remains unknown. Commonly, 37°C is used to culture embryos and is based on the estimate of human core body temperature. However, body temperature varies and is likely not exactly 37°C. Temperature can vary based on time of day, sex, and between individuals. Furthermore, temperature of the female reproductive tract may actually be slightly less than core body temperature. The temperature of the human follicle is approximately 2.3°C cooler than core body temperature, and in animal models the fallopian tube is approximately 1.5°C cooler than core body temperature.133–138 As a result, the question has been posed “whether human IVF and related procedures should be carried out, at, say, 35.5 to 36°C rather than at 37°C.”130 Use of a 1.5°C lower temperature would presumably lower the metabolic rate of the embryo approximately 15%, which may result in a more “quiet” metabolism, thereby possibly benefiting embryo development.130 However, this hypothesis has not been proven. Interestingly, retrospective, observational data on incubator performance indicated that an increased pregnancy rate was obtained from incubators with a temperature of 36.96° C  0.13 compared with those with a temperature of 37.03° C  0.13.122 However, this study was not well controlled and results cannot be attributed to temperature alone. In addition, biologic explanation for how such a small temperature difference (0.07), which falls into the acceptable variation range of many thermometers, could result in observed developmental differences in unknown. However, this has led some

Table 6 Effects of reduced oxygen concentration in a predominantly blastocyst transfer program Endpoint

21% O2

5% O2

p value

Clinical pregnancy rate

56/115 (48.7%)

74/115 (64.3%)

0.027

Implantation

95/267 (35.6%)

122/247 (49.4%

0.003

Live birth

49/115 (42.6%)

66/115 (57.4%)

0.043

Note: When examining all patients in a prospective randomized trial, extended culture in low oxygen significantly improved clinical pregnancy, implantation, and live birth. Source: Adapted from Meintjes et al.70,113 Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

109

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

Optimal Human Embryo Culture

Optimal Human Embryo Culture

Swain

laboratories to culture embryos around 36.5 to 36.8°C in order to avoid the possibility of temperature rising above 37°C. It has been shown that overheating oocytes to a temperature of 39°C or greater is detrimental to formation of the meiotic spindle during maturation in vitro.127 However, there is no proof that 37°C is too high for human embryo culture or that this may be detrimental. In fact, animal embryos have mechanisms that help convey thermotolerance to deal with slight elevations in temperature,139–142 and the same may be true for human. Recently, a prospective randomized controlled trial utilized patient embryo splits and examined the effect of culture at either 37 or 36°C.143 Controlling for temperature variations, incubator type, and pH, the study demonstrated that culture of human embryos at 37°C yielded higher average cell numbers at day 3 of development, higher blastocyst formation, and higher useable blastocyst rates compared with 36°C. No differences were observed in rates of aneuploidy or implantation (►Fig. 1). More detailed analysis of the optimal temperature for IVF is required, but the best practice at the moment appears to be maintaining temperature around 37°C. There is no clear advantage of one incubator type over another in terms of temperature regulation, though some types may recover temperature faster or maintain temperature better than others under various circumstances.98

Group Culture Embryo density: It is known that embryos secrete various proteins during their culture, and research analyzing the

embryo secretome has become an active area of research.144–148 In addition, using sensitive chemical microprobes, gradients of dissolved oxygen, potassium, and calcium have been shown to extend around the embryo(s) that are measurable beyond 50µm from mouse embryos cultured in vitro.149–151 Thus, embryos may modify their own microenvironment and this may impact embryo development. Considering the ability of embryos to modify their microenvironment, group culture has long been known to be beneficial for animal embryos, from both mono- and polyovulatory species. Various studies have shown a benefit of group culture of embryos from the rodent152–156 as well as domestic species such as the cow, sheep, and pig.157–164 Various factors involved in group embryo culture are likely important, and they include the ratio of media volume used to the number of embryos (embryo density), embryo spacing, as well as quality of the companion embryos.165 While detailed analysis on factors such as embryo density, spacing, or identification of zones created around human embryos is largely lacking, there are studies indicating beneficial effects of group embryo culture. Group culture of embryos significantly enhanced cleavage rates but not embryo morphology grades as compared with embryos grown individually in one study of 55 patients.166 Another small study found significantly improved clinical pregnancy rates with group culture as compared with individually cultured embryos (43 vs. 24%).167 Rebollar-Lazaro and Matson168 retrospectively analyzed patients who had embryos cultured from days 1 to 3, either individually or in groups of three to five embryos. Culturing in groups for the first 3 days had no impact on pregnancy or implantation rates compared with

Fig. 1 Impact of different culture temperature on human embryo development. (a) A sibling embryo split design was used, and incubator type, temperature variation and pH were accounted for as confounding variables. (b) Data indicate that 37°C is superior to 36°C in terms of embryo development. Different superscripts within a column represent statistically significant differences between temperature treatments p < 0.05 (Adapted from Hong et al, 2014.) Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

110

Optimal Human Embryo Culture

Swain

111

Table 7 Impact of group culture on human embryo development Individual culture Compaction

98 (38.9%)

a a

Group culture with no contact

Group culture with contact

116 (38.2%)

a

188 (49.5%)b

124 (40.8%)

a

212 (55.8%)b

Blastocyst formation

114 (42.5%)

Blastocyst quality

44/68 (64.7%)a

66/96 (68.6%)a,b

122/154 (79.2%)b

Live birth

5 (38.5%)

5 (41.7%)

28 (62.2%)

those embryos cultured individually, but did result in more usable blastocysts. This is similar to a prospective report that indicated no benefit of group culture of human embryos prior to day 2 or 3 transfer,169 suggesting extended culture may be required to observe the full benefits of communal culture. Finally, an interesting prospective trial of 72 patients compared outcomes of embryos split between three treatment groups: individually cultured embryos, embryos cultured in groups in which embryos were physically separated but able to exchange media, or embryos cultured in groups without physical separation. Importantly, all treatments were included within the same culture dish to help control for variables. Group culture of embryos without physical separation resulted in greater rates of compaction on day 4 and blastocyst formation on day 5 compared with group culture with physical separation and embryos cultured individually. There was a trend toward a higher live birth rate with group culture of embryos without physical separation compared with individual culture (►Table 7). This indicates that extended group culture of embryos may be beneficial for human preimplantation embryo development, and that physical contact or proximity of embryos seems to be an important factor.

Culture Dishes Importantly, new micro- or nanovolume culture dishes are being developed, and use of these dishes to optimize the microenvironment may help further improve the benefits of group embryo culture, as discussed previously. Additionally, these dishes could perhaps improve individual embryo culture or be used for study of embryo metabolomics, time-lapse imaging, or embryo tracking for genetic diagnosis.171–173 At least one preliminary study indicates that a microvolume approach using the well-of-a-well (WOW) dish may be superior for human embryo development compared with larger volume culture.174 This WOW approach uses microwells to house individual embryos, but all embryos share a common media overlay and reside in close proximity to each other. Currently, there is no consensus on a superior culture dish or volume of media used for culture. Directly related to optimizing the culture system is the need to ensure that the culture dish and other contact materials used are nontoxic. Though the importance of oil and avoiding toxicity has been discussed, this also includes culture dishes and other contact materials that could

inadvertently alter media characteristics and growth conditions to impair embryo development. Contact material testing is one of the most important aspects of laboratory QC/QA and is paramount for optimizing the culture system. It is well documented that not all products, despite being packaged or sold as sterile, are inert in terms of impact on embryo development.175,176 Thus, verification of material safety is required prior to clinical use. This is usually performed using a relevant bioassay, such as the mouse embryo assay (MEA) or the human sperm survival assay (HSSA), also known as the human sperm motility assay (HSMA). Comparisons and the merits of these two assays have been discussed elsewhere177–183 and each can be useful in their own respect, with factors such as cost and availability warranting attention. Perhaps more relevant, the sensitivity of the bioassay is important and approaches can be modified to increase this sensitivity to ensure that subtle material toxicity can be detected. Examples of approaches used to increase sensitivity are use of the one- versus the two-cell MEA, outbred versus inbred versus hybrid mouse strains, blastocyst cell counts versus simple blastocyst formation, exclusion of protein from media versus protein inclusion, and use of a simple media versus a more robust complex media.184–186 Importantly, thresholds must be set for an assay to “pass,” and these thresholds should be rigorous. Each laboratory can determine which assay and threshold suits their needs regarding sensitivity and cost. A commonly used and often recommended sensitive assay includes using the one-cell MEA, noting time-appropriate embryo development with rate of expanded or hatching blastocyst of greater than70% at 96 hours (4 days) of culture. Real-time assessment of morphokinetics may also be useful in helping improve sensitivity of the MEA.187 One interesting point is that many materials from manufacturers can now be purchased as “pretested,” having passed some manner of bioassay. However, these manufacturer’s bioassays may not meet the sensitivity criteria of individual laboratories. Furthermore, factors impacting material quality can convey detrimental effects at a point following initial successful passing of the bioassay. For example, improper storage conditions in warehouses or delayed effects from prolonged storage can cause a “prescreened” item to later become toxic once received in the laboratory. A common example of this appears to be toxicity associated with mineral oil.73,175,187 Also, potential toxicity from a single item could be very minor, but when used in the context of several Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

Note: Use of sibling embryos indicates that group culture improves embryo development following extended culture. Different superscripts represent statistically significant differences between culture treatments, p < 0.05. Source: Adapted from Ebner et al.170

Optimal Human Embryo Culture

Swain

sequential items used in a particular laboratory culture system, this toxicity could be compounded due to interactions with other system components. This has been noted anecdotally across several laboratories, in which “tested” materials are later found to have a detrimental impact. Therefore, more stringent laboratories often retest certain prescreened materials, testing them under their own laboratory conditions before releasing them for clinical use.

4 Ebner T, Moser M, Shebl O, Sommergruber M, Yaman C, Tews G.

Conclusion

7

It is evident that there are numerous variables to consider when optimizing an embryo culture system. Though most variables are consistent between laboratories, small differences in laboratory conditions will affect what is the “optimal” embryo culture system for a given program. Of course, even if a culture system is optimized, embryo selection remains a critically important factor to maximize outcomes. In fact, if a system is optimized to maximize embryo development, selection becomes even more difficult, as there are more “good-quality” embryos from which to choose. The best embryo selection method to use then becomes critically important and is the topic of much active research. That being said, the following key, evidenced-based factors should be included in a highly successful culture system: 1. Human eggs should be cultured for 3 to 4 hours with cumulus cells intact following retrieval to allow for optimal egg maturation prior to ICSI. 2. Use of low oxygen should be followed for all embryo culture, with the most benefit occurring when used for extended culture to the blastocyst stage. 3. Careful attention to detail is required when preparing culture dishes to avoid increased in media osmolality, which can compromise embryo development. 4. Culture media should be selected based on an individual laboratories performance, but media should include some assortment of amino acids, and if containing glutamine, should contain a dipeptide form. 5. Careful control of incubator workflow is required to maximize embryo development, making sure that sufficient numbers of incubators are available to help limit openings and closings to keep strict control of atmosphere (pH) and temperature. 6. Maternal testing and an appropriate biossay should be in place to avoid toxicity.

5

6

8

9

10

11

12

13

14

15

16

17

18

19

References 1 Hyun CS, Cha JH, Son WY, Yoon SH, Kim KA, Lim JH. Optimal ICSI

timing after the first polar body extrusion in in vitro matured human oocytes. Hum Reprod 2007;22(7):1991–1995 2 Rienzi L, Ubaldi F, Anniballo R, Cerulo G, Greco E. Preincubation of human oocytes may improve fertilization and embryo quality after intracytoplasmic sperm injection. Hum Reprod 1998;13(4): 1014–1019 3 Hassan HA. Cumulus cell contribution to cytoplasmic maturation and oocyte developmental competence in vitro. J Assist Reprod Genet 2001;18(10):539–543

Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

20

21

Blood clots in the cumulus-oocyte complex predict poor oocyte quality and post-fertilization development. Reprod Biomed Online 2008;16(6):801–807 Daya S, Kohut J, Gunby J, Younglai E. Influence of blood clots in the cumulus complex on oocyte fertilization and cleavage. Hum Reprod 1990;5(6):744–746 Van de Velde H, De Vos A, Joris H, Nagy ZP, Van Steirteghem AC. Effect of timing of oocyte denudation and micro-injection on survival, fertilization and embryo quality after intracytoplasmic sperm injection. Hum Reprod 1998;13(11):3160–3164 Rienzi L, Ubaldi F, Martinez F, et al. Relationship between meiotic spindle location with regard to the polar body position and oocyte developmental potential after ICSI. Hum Reprod 2003; 18(6):1289–1293 Taylor TH, Wright G, Jones-Colon S, Mitchell-Leef D, Kort HI, Nagy ZP. Comparison of ICSI and conventional IVF in patients with increased oocyte immaturity. Reprod Biomed Online 2008;17(1):46–52 Evison M, Pretty C, Taylor E, Franklin C. Human recombinant hyaluronidase (Cumulase) improves intracytoplasmic sperm injection survival and fertilization rates. Reprod Biomed Online 2009;18(6):811–814 Taylor TH, Elliott T, Colturato LF, Straub RJ, Mitchell-Leef D, Nagy ZP. Comparison of bovine- and recombinant human-derived hyaluronidase with regard to fertilization rates and embryo morphology in a sibling oocyte model: a prospective, blinded, randomized study. Fertil Steril 2006;85(5):1544–1546 De Vos A, Van Landuyt L, Van Ranst H, et al. Randomized siblingoocyte study using recombinant human hyaluronidase versus bovine-derived Sigma hyaluronidase in ICSI patients. Hum Reprod 2008;23(8):1815–1819 Sakkas D. Novel technologies for selecting the best sperm for in vitro fertilization and intracytoplasmic sperm injection. Fertil Steril 2013;99(4):1023–1029 Said TM, Land JA. Effects of advanced selection methods on sperm quality and ART outcome: a systematic review. Hum Reprod Update 2011;17(6):719–733 Jayaraman V, Upadhya D, Narayan PK, Adiga SK. Sperm processing by swim-up and density gradient is effective in elimination of sperm with DNA damage. J Assist Reprod Genet 2012;29(6):557–563 Teixeira DM, Barbosa MA, Ferriani RA, et al. Regular (ICSI) versus ultra-high magnification (IMSI) sperm selection for assisted reproduction. Cochrane Database Syst Rev 2013;7:CD010167 Leandri RD, Gachet A, Pfeffer J, et al. Is intracytoplasmic morphologically selected sperm injection (IMSI) beneficial in the first ART cycle? A multicentric randomized controlled trial. Andrology 2013;1(5):692–697 De Vos A, Van de Velde H, Bocken G, et al. Does intracytoplasmic morphologically selected sperm injection improve embryo development? A randomized sibling-oocyte study. Hum Reprod 2013;28(3):617–626 Delaroche L, Yazbeck C, Gout C, Kahn V, Oger P, Rougier N. Intracytoplasmic morphologically selected sperm injection (IMSI) after repeated IVF or ICSI failures: a prospective comparative study. Eur J Obstet Gynecol Reprod Biol 2013;167(1):76–80 Worrilow KC, Eid S, Woodhouse D, et al. Use of hyaluronan in the selection of sperm for intracytoplasmic sperm injection (ICSI): significant improvement in clinical outcomes—multicenter, double-blinded and randomized controlled trial. Hum Reprod 2013; 28(2):306–314 Majumdar G, Majumdar A. A prospective randomized study to evaluate the effect of hyaluronic acid sperm selection on the intracytoplasmic sperm injection outcome of patients with unexplained infertility having normal semen parameters. J Assist Reprod Genet 2013;30(11):1471–1475 Pool TB, Schoolfield J, Han D. Human embryo culture media comparisons. Methods Mol Biol 2012;912:367–386

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

112

Optimal Human Embryo Culture

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

and IVF/ICSI success rates: a systematic review. Hum Reprod Update 2013;19(3):210–220 Mauri AL, Petersen CG, Baruffi RL, Franco JG Jr. A prospective, randomized comparison of two commercial media for ICSI and embryo culture. J Assist Reprod Genet 2001;18(7):378–381 Ben-Yosef D, Amit A, Azem F, et al. Prospective randomized comparison of two embryo culture systems: P1 medium by Irvine Scientific and the Cook IVF Medium. J Assist Reprod Genet 2004; 21(8):291–295 Van Langendonckt A, Demylle D, Wyns C, Nisolle M, Donnez J. Comparison of G1.2/G2.2 and Sydney IVF cleavage/blastocyst sequential media for the culture of human embryos: a prospective, randomized, comparative study. Fertil Steril 2001;76(5): 1023–1031 Balaban B, Urman B. Comparison of two sequential media for culturing cleavage-stage embryos and blastocysts: embryo characteristics and clinical outcome. Reprod Biomed Online 2005; 10(4):485–491 Zollner KP, Zollner U, Schneider M, Dietl J, Steck T. Comparison of two media for sequential culture after IVF and ICSI shows no differences in pregnancy rates: a randomized trial. Med Sci Monit 2004;10(1):CR1–CR7 Hambiliki F, Sandell P, Yaldir F, Stavreus-Evers A. A prospective randomized sibling-oocyte study of two media systems for culturing cleavage-stage embryos-impact on fertilization rate. J Assist Reprod Genet 2011;28(4):335–341 Sifer C, Handelsman D, Grange E, et al. An auto-controlled prospective comparison of two embryos culture media (G III series versus ISM) for IVF and ICSI treatments. J Assist Reprod Genet 2009;26(11-12):575–581 Paternot G, Debrock S, D’Hooghe TM, Spiessens C. Early embryo development in a sequential versus single medium: a randomized study. Reprod Biol Endocrinol 2010;8:83 Sepúlveda S, Garcia J, Arriaga E, Diaz J, Noriega-Portella L, Noriega-Hoces L. In vitro development and pregnancy outcomes for human embryos cultured in either a single medium or in a sequential media system. Fertil Steril 2009;91(5):1765–1770 Basile N, Morbeck D, García-Velasco J, Bronet F, Meseguer M. Type of culture media does not affect embryo kinetics: a timelapse analysis of sibling oocytes. Hum Reprod 2013;28(3): 634–641 Ciray HN, Aksoy T, Goktas C, Ozturk B, Bahceci M. Time-lapse evaluation of human embryo development in single versus sequential culture media—a sibling oocyte study. J Assist Reprod Genet 2012;29(9):891–900 Reed ML, Hamic A, Thompson DJ, Caperton CL. Continuous uninterrupted single medium culture without medium renewal versus sequential media culture: a sibling embryo study. Fertil Steril 2009;92(5):1783–1786 Wales RG, Whittingham DG. Decomposition of sodium pyruvate in culture media stored at 5 degrees C and its effects on the development of the preimplantation mouse embryo. J Reprod Fertil 1971;24(1):126 Stewart-Savage J, Bavister BD. Deterioration of stored culture media as monitored by a sperm motility bioassay. J In Vitro Fert Embryo Transf 1988;5(2):76–80 Edwards LJ, Williams DA, Gardner DK. Intracellular pH of the preimplantation mouse embryo: effects of extracellular pH and weak acids. Mol Reprod Dev 1998;50(4):434–442 Carrillo AJ, Lane B, Pridman DD, et al. Improved clinical outcomes for in vitro fertilization with delay of embryo transfer from 48 to 72 hours after oocyte retrieval: use of glucose- and phosphatefree media. Fertil Steril 1998;69(2):329–334 Quinn P, Moinipanah R, Steinberg JM, Weathersbee PS. Successful human in vitro fertilization using a modified human tubal fluid medium lacking glucose and phosphate ions. Fertil Steril 1995; 63(4):922–924

113

40 Lane M. Mechanisms for managing cellular and homeostatic

stress in vitro. Theriogenology 2001;55(1):225–236 41 Gardner DK, Lane M. Alleviation of the ‘2-cell block’ and devel-

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59 60

opment to the blastocyst of CF1 mouse embryos: role of amino acids, EDTA and physical parameters. Hum Reprod 1996;11(12): 2703–2712 McKiernan SH, Clayton MK, Bavister BD. Analysis of stimulatory and inhibitory amino acids for development of hamster one-cell embryos in vitro. Mol Reprod Dev 1995;42(2):188–199 Lane M, Gardner DK. Differential regulation of mouse embryo development and viability by amino acids. J Reprod Fertil 1997; 109(1):153–164 Lane M, Hooper K, Gardner DK. Effect of essential amino acids on mouse embryo viability and ammonium production. J Assist Reprod Genet 2001;18(9):519–525 Devreker F, Winston RM, Hardy K. Glutamine improves human preimplantation development in vitro. Fertil Steril 1998;69(2): 293–299 Devreker F, Van den Bergh M, Biramane J, Winston RL, Englert Y, Hardy K. Effects of taurine on human embryo development in vitro. Hum Reprod 1999;14(9):2350–2356 Hammer MA, Kolajova M, Léveillé M, Claman P, Baltz JM. Glycine transport by single human and mouse embryos. Hum Reprod 2000;15(2):419–426 Biggers JD, McGinnis LK, Lawitts JA. Enhanced effect of glycyl-Lglutamine on mouse preimplantation embryos in vitro. Reprod Biomed Online 2004;9(1):59–69 Summers MC, McGinnis LK, Lawitts JA, Biggers JD. Mouse embryo development following IVF in media containing either L-glutamine or glycyl-L-glutamine. Hum Reprod 2005;20(5):1364–1371 Lane M, Gardner DK. Ammonium induces aberrant blastocyst differentiation, metabolism, pH regulation, gene expression and subsequently alters fetal development in the mouse. Biol Reprod 2003;69(4):1109–1117 Virant-Klun I, Tomazevic T, Vrtacnik-Bokal E, Vogler A, Krsnik M, Meden-Vrtovec H. Increased ammonium in culture medium reduces the development of human embryos to the blastocyst stage. Fertil Steril 2006;85(2):526–528 Kastrop PM, de Graaf-Miltenburg LA, Gutknecht DR, Weima SM. Microbial contamination of embryo cultures in an ART laboratory: sources and management. Hum Reprod 2007;22(8): 2243–2248 Magli MC, Gianaroli L, Fiorentino A, Ferraretti AP, Fortini D, Panzella S. Improved cleavage rate of human embryos cultured in antibiotic-free medium. Hum Reprod 1996;11(7):1520–1524 Zhou H, McKiernan SH, Ji W, Bavister BD. Effect of antibiotics on development in vitro of hamster pronucleate ova. Theriogenology 2000;54(7):999–1006 Liu J, Tang S, Xu W, Wang Y, Yin B, Zhang Y. Detrimental effects of antibiotics on mouse embryos in chromatin integrity, apoptosis and expression of zygotically activated genes. Zygote 2011;19(2): 137–145 Zhu J, Li M, Chen L, Liu P, Qiao J. The protein source in embryo culture media influences birthweight: a comparative study between G1 v5 and G1-PLUS v5. Hum Reprod 2014;29(7):1387–1392 Leonard PH, Charlesworth MC, Benson L, Walker DL, Fredrickson JR, Morbeck DE. Variability in protein quality used for embryo culture: embryotoxicity of the stabilizer octanoic acid. Fertil Steril 2013;100(2):544–549 Morbeck DE, Paczkowski M, Fredrickson JR, et al. Composition of protein supplements used for human embryo culture. J Assist Reprod Genet 2014;31(12):1703–1711 Meintjes M. Media composition: macromolecules and embryo growth. Methods Mol Biol 2012;912:107–127 Dyrlund TF, Kirkegaard K, Poulsen ET, et al. Unconditioned commercial embryo culture media contain a large variety of non-declared proteins: a comprehensive proteomics analysis. Hum Reprod 2014;29(11):2421–2430 Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

22 Mantikou E, Youssef MA, van Wely M, et al. Embryo culture media

Swain

Optimal Human Embryo Culture

Swain

61 Borini A, Bulletti C, Cattoli M, et al. Use of recombinant leukemia

62

63

64 65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

inhibitory factor in embryo implantation. Ann N Y Acad Sci 1997; 828:157–161 Sjöblom C, Wikland M, Robertson SA. Granulocyte-macrophage colony-stimulating factor promotes human blastocyst development in vitro. Hum Reprod 1999;14(12):3069–3076 O’Neill C, Ryan JP, Collier M, Saunders DM, Ammit AJ, Pike IL. Supplementation of in-vitro fertilisation culture medium with platelet activating factor. Lancet 1989;2(8666):769–772 Hegde A, Behr B. Media composition: growth factors. Methods Mol Biol 2012;912:177–198 Weathersbee PS, Pool TB, Ord T. Synthetic serum substitute (SSS): a globulin-enriched protein supplement for human embryo culture. J Assist Reprod Genet 1995;12(6):354–360 Pool TB, Martin JE. High continuing pregnancy rates after in vitro fertilization-embryo transfer using medium supplemented with a plasma protein fraction containing alpha- and beta-globulins. Fertil Steril 1994;61(4):714–719 Schneider EG, Hayslip CC. Globulin-enriched protein supplements shorten the pre-compaction mitotic interval and promote hatching of murine embryos. Am J Reprod Immunol 1996;36(2):101–106 Desai NN, Sheean LA, Martin D, et al. Clinical experience with synthetic serum substitute as a protein supplement in IVF culture media: a retrospective study. J Assist Reprod Genet 1996;13(1): 23–31 Tucker KE, Hurst BS, Guadagnoli S, et al. Evaluation of synthetic serum substitute versus serum as protein supplementation for mouse and human embryo culture. J Assist Reprod Genet 1996; 13(1):32–37 Meintjes M, Chantilis SJ, Ward DC, et al. A randomized controlled study of human serum albumin and serum substitute supplement as protein supplements for IVF culture and the effect on live birth rates. Hum Reprod 2009;24(4):782–789 Bungum M, Humaidan P, Bungum L. Recombinant human albumin as protein source in culture media used for IVF: a prospective randomized study. Reprod Biomed Online 2002;4(3):233–236 Ben-Yosef D, Yovel I, Schwartz T, Azem F, Lessing JB, Amit A. Increasing synthetic serum substitute (SSS) concentrations in P1 glucose/phosphate-free medium improves implantation rate: a comparative study. J Assist Reprod Genet 2001;18(11):588–592 Morbeck DE, Khan Z, Barnidge DR, Walker DL. Washing mineral oil reduces contaminants and embryotoxicity. Fertil Steril 2010; 94(7):2747–2752 Otsuki J, Nagai Y, Chiba K. Damage of embryo development caused by peroxidized mineral oil and its association with albumin in culture. Fertil Steril 2009;91(5):1745–1749 Otsuki J, Nagai Y, Chiba K. Peroxidation of mineral oil used in droplet culture is detrimental to fertilization and embryo development. Fertil Steril 2007;88(3):741–743 Hughes PM, Morbeck DE, Hudson SB, Fredrickson JR, Walker DL, Coddington CC. Peroxides in mineral oil used for in vitro fertilization: defining limits of standard quality control assays. J Assist Reprod Genet 2010;27(2–3):87–92 Lane M, Lyons EA, Bavister BD. Cryopreservation reduces the ability of hamster 2-cell embryos to regulate intracellular pH. Hum Reprod 2000;15(2):389–394 Zander-Fox DL, Mitchell M, Thompson JG, Lane M. Alterations in mouse embryo intracellular pH by DMO during culture impair implantation and fetal growth. Reprod Biomed Online 2010; 21(2):219–229 Squirrell JM, Lane M, Bavister BD. Altering intracellular pH disrupts development and cellular organization in preimplantation hamster embryos. Biol Reprod 2001;64(6):1845–1854 Phillips KP, Léveillé MC, Claman P, Baltz JM. Intracellular pH regulation in human preimplantation embryos. Hum Reprod 2000;15(4):896–904 Phillips KP, Petrunewich MA, Collins JL, Baltz JM. The intracellular pH-regulatory HCO3-/Cl- exchanger in the mouse oocyte is

Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

82 83 84 85

86

87

88

89

90

91 92

93

94

95

96

97

98

99

100

101

inactivated during first meiotic metaphase and reactivated after egg activation via the MAP kinase pathway. Mol Biol Cell 2002; 13(11):3800–3810 Quinn P. Culture systems: sequential. Methods Mol Biol 2012; 912:211–230 Hentemann M, Mousavi K, Bertheussen K. Differential pH in embryo culture. Fertil Steril 2011;95(4):1291–1294 Swain JE. Is there an optimal pH for culture media used in clinical IVF? Hum Reprod Update 2012;18(3):333–339 Moreno-Cuevas JE, Sirbasku DA. Estrogen mitogenic action. III. is phenol red a “red herring”? In Vitro Cell Dev Biol Anim 2000; 36(7):447–464 Nakayama T, Noda Y, Goto Y, Mori T. Effects of visible light and other environmental factors on the production of oxygen radicals by hamster embryos. Theriogenology 1994;41(2): 499–510 Morgia F, Torti M, Montigiani M, et al. Use of a medium buffered with N-hydroxyethylpiperazine-N-ethanesulfonate (HEPES) in intracytoplasmic sperm injection procedures is detrimental to the outcome of in vitro fertilization. Fertil Steril 2006;85(5): 1415–1419 Swain JE, Pool TB. New pH-buffering system for media utilized during gamete and embryo manipulations for assisted reproduction. Reprod Biomed Online 2009;18(6):799–810 Will MA, Clark NA, Swain JE. Biological pH buffers in IVF: help or hindrance to success. J Assist Reprod Genet 2011;28(8): 711–724 Swain JE. Optimizing the culture environment in the IVF laboratory: impact of pH and buffer capacity on gamete and embryo quality. Reprod Biomed Online 2010;21(1):6–16 Swain JE. Media composition: pH and buffers. Methods Mol Biol 2012;912:161–175 Hadi T, Hammer MA, Algire C, Richards T, Baltz JM. Similar effects of osmolarity, glucose, and phosphate on cleavage past the 2-cell stage in mouse embryos from outbred and F1 hybrid females. Biol Reprod 2005;72(1):179–187 Baltz JM, Tartia AP. Cell volume regulation in oocytes and early embryos: connecting physiology to successful culture media. Hum Reprod Update 2010;16(2):166–176 Dawson KM, Baltz JM. Organic osmolytes and embryos: substrates of the Gly and beta transport systems protect mouse zygotes against the effects of raised osmolarity. Biol Reprod 1997; 56(6):1550–1558 Dawson KM, Collins JL, Baltz JM. Osmolarity-dependent glycine accumulation indicates a role for glycine as an organic osmolyte in early preimplantation mouse embryos. Biol Reprod 1998; 59(2):225–232 Richards T, Wang F, Liu L, Baltz JM. Rescue of postcompactionstage mouse embryo development from hypertonicity by amino acid transporter substrates that may function as organic osmolytes. Biol Reprod 2010;82(4):769–777 Swain JE, Cabrera L, Xu X, Smith GD. Microdrop preparation factors influence culture-media osmolality, which can impair mouse embryo preimplantation development. Reprod Biomed Online 2012;24(2):142–147 Swain JE. Decisions for the IVF laboratory: comparative analysis of embryo culture incubators. Reprod Biomed Online 2014;28(5): 535–547 Hyslop L, Prathalingam N, Nowak L, et al. A novel isolator-based system promotes viability of human embryos during laboratory processing. PLoS ONE 2012;7(2):e31010 Zhang JQ, Li XL, Peng Y, Guo X, Heng BC, Tong GQ. Reduction in exposure of human embryos outside the incubator enhances embryo quality and blastulation rate. Reprod Biomed Online 2010;20(4):510–515 Kea B, Gebhardt J, Watt J, et al. Effect of reduced oxygen concentrations on the outcome of in vitro fertilization. Fertil Steril 2007; 87(1):213–216

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

114

Swain

102 Dumoulin JC, Meijers CJ, Bras M, Coonen E, Geraedts JP, Evers JL.

122 Higdon HL III, Blackhurst DW, Boone WR. Incubator management

Effect of oxygen concentration on human in-vitro fertilization and embryo culture. Hum Reprod 1999;14(2):465–469 Catt JW, Henman M. Toxic effects of oxygen on human embryo development. Hum Reprod 2000;15(Suppl 2):199–206 Ciray HN, Aksoy T, Yaramanci K, Karayaka I, Bahceci M. In vitro culture under physiologic oxygen concentration improves blastocyst yield and quality: a prospective randomized survey on sibling oocytes. Fertil Steril 2009;91(4, Suppl):1459–1461 Bahçeci M, Ciray HN, Karagenc L, Uluğ U, Bener F. Effect of oxygen concentration during the incubation of embryos of women undergoing ICSI and embryo transfer: a prospective randomized study. Reprod Biomed Online 2005;11(4):438–443 Kovacic B, Sajko MC, Vlaisavljević V. A prospective, randomized trial on the effect of atmospheric versus reduced oxygen concentration on the outcome of intracytoplasmic sperm injection cycles. Fertil Steril 2010;94(2):511–519 Kovacic B, Vlaisavljević V. Influence of atmospheric versus reduced oxygen concentration on development of human blastocysts in vitro: a prospective study on sibling oocytes. Reprod Biomed Online 2008;17(2):229–236 Nanassy L, Peterson CA, Wilcox AL, Peterson CM, Hammoud A, Carrell DT. Comparison of 5% and ambient oxygen during days 3-5 of in vitro culture of human embryos. Fertil Steril 2010;93(2): 579–585 Waldenström U, Engström AB, Hellberg D, Nilsson S. Low-oxygen compared with high-oxygen atmosphere in blastocyst culture, a prospective randomized study. Fertil Steril 2009;91(6):2461–2465 Gomes Sobrinho DB, Oliveira JB, Petersen CG, et al. IVF/ICSI outcomes after culture of human embryos at low oxygen tension: a meta-analysis. Reprod Biol Endocrinol 2011;9:143 Petersen A, Mikkelsen AL, Lindenberg S. The impact of oxygen tension on developmental competence of post-thaw human embryos. Acta Obstet Gynecol Scand 2005;84(12):1181–1184 Bedaiwy MA, Mahfouz RZ, Goldberg JM, et al. Relationship of reactive oxygen species levels in day 3 culture media to the outcome of in vitro fertilization/intracytoplasmic sperm injection cycles. Fertil Steril 2010;94(6):2037–2042 Meintjes M, Chantilis SJ, Douglas JD, et al. A controlled randomized trial evaluating the effect of lowered incubator oxygen tension on live births in a predominantly blastocyst transfer program. Hum Reprod 2009;24(2):300–307 Kasterstein E, Strassburger D, Komarovsky D, et al. The effect of two distinct levels of oxygen concentration on embryo development in a sibling oocyte study. J Assist Reprod Genet 2013;30(8): 1073–1079 Thomas T. Culture systems: air quality. Methods Mol Biol 2012; 912:313–324 Hall J, Gilligan A, Schimmel T, Cecchi M, Cohen J. The origin, effects and control of air pollution in laboratories used for human embryo culture. Hum Reprod 1998;13(Suppl 4):146–155 Cohen J, Gilligan A, Willadsen S. Culture and quality control of embryos. Hum Reprod 1998;13(Suppl 3):137–144, discussion 145–147 Cohen J, Gilligan A, Esposito W, Schimmel T, Dale B. Ambient air and its potential effects on conception in vitro. Hum Reprod 1997; 12(8):1742–1749 Boone WR, Johnson JE, Locke AJ, Crane MM IV, Price TM. Control of air quality in an assisted reproductive technology laboratory. Fertil Steril 1999;71(1):150–154 Khoudja RY, Xu Y, Li T, Zhou C. Better IVF outcomes following improvements in laboratory air quality. J Assist Reprod Genet 2013;30(1):69–76 Merton JS, Vermeulen ZL, Otter T, Mullaart E, de Ruigh L, Hasler JF. Carbon-activated gas filtration during in vitro culture increased pregnancy rate following transfer of in vitro-produced bovine embryos. Theriogenology 2007;67(7):1233–1238

in an assisted reproductive technology laboratory. Fertil Steril 2008;89(3):703–710 Schimmel T, Gilligan A, Garrisi G, et al. Removal of volatile organic compounds from incubators used for gamete and embryo culture. Fertil Steril 1997;68(Suppl 1):s165 Mayer J, Nechchiri F, Weedon V, et al. Prospective randomized crossover analysis of the impact of an IVF incubator air filtration system (coda, GenX) on clinical pregnancy rates. Fertil Steril Suppl 1999;72(Suppl 1):S42 McLellan S, Panagoulopoulos C, Dickinson K, et al. Effect of incubator air filtration system on IVF outcomes. Fertil Steril 2001;76(Suppl 1):s103 Battaglia D, Khabani A, Rainer C, et al. Prospective randomized trial of incubator CODA filtration unites revealed no effect on outcome. Fertil Steril 2001;75(Suppl 1):s6 Sun XF, Wang WH, Keefe DL. Overheating is detrimental to meiotic spindles within in vitro matured human oocytes. Zygote 2004;12(1):65–70 Wang WH, Meng L, Hackett RJ, Oldenbourg R, Keefe DL. Rigorous thermal control during intracytoplasmic sperm injection stabilizes the meiotic spindle and improves fertilization and pregnancy rates. Fertil Steril 2002;77(6):1274–1277 Wang WH, Meng L, Hackett RJ, Odenbourg R, Keefe DL. Limited recovery of meiotic spindles in living human oocytes after cooling-rewarming observed using polarized light microscopy. Hum Reprod 2001;16(11):2374–2378 Leese HJ, Baumann CG, Brison DR, McEvoy TG, Sturmey RG. Metabolism of the viable mammalian embryo: quietness revisited. Mol Hum Reprod 2008;14(12):667–672 Pollard JW, Martino A, Rumph ND, Songsasen N, Plante C, Leibo SP. Effect of ambient temperatures during oocyte recovery on in vitro production of bovine embryos. Theriogenology 1996;46(5): 849–858 Lane M, Mitchell M, Cashman KS, Feil D, Wakefield S, Zander-Fox DL. To QC or not to QC: the key to a consistent laboratory? Reprod Fertil Dev 2008;20(1):23–32 Grinsted J, Kjer JJ, Blendstrup K, Pedersen JF. Is low temperature of the follicular fluid prior to ovulation necessary for normal oocyte development? Fertil Steril 1985;43(1):34–39 David A, Vilensky A, Nathan H. Temperature changes in different parts of the rabbit oviduct. Preliminary report [in Hebrew]. Harefuah 1971;80(4):180–182 Hunter RH, Bogh IB, Einer-Jensen N, Müller S, Greve T. Preovulatory graafian follicles are cooler than neighbouring stroma in pig ovaries. Hum Reprod 2000;15(2):273–283 Hunter RH, Grøndahl C, Greve T, Schmidt M. Graafian follicles are cooler than neighbouring ovarian tissues and deep rectal temperatures. Hum Reprod 1997;12(1):95–100 Hunter RH, Nichol R. A preovulatory temperature gradient between the isthmus and ampulla of pig oviducts during the phase of sperm storage. J Reprod Fertil 1986;77(2):599–606 Hunter RH. Temperature gradients in female reproductive tissues. Reprod Biomed Online 2012;24(4):377–380 Edwards JL, Hansen PJ. Elevated temperature increases heat shock protein 70 synthesis in bovine two-cell embryos and compromises function of maturing oocytes. Biol Reprod 1996;55(2): 341–346 Edwards JL, Hansen PJ. Differential responses of bovine oocytes and preimplantation embryos to heat shock. Mol Reprod Dev 1997;46(2):138–145 Chandolia RK, Peltier MR, Tian W, Hansen PJ. Transcriptional control of development, protein synthesis, and heat-induced heat shock protein 70 synthesis in 2-cell bovine embryos. Biol Reprod 1999;61(6):1644–1648 Aréchiga CF, Hansen PJ. Response of preimplantation murine embryos to heat shock as modified by developmental stage

103 104

105

106

107

108

109

110

111

112

113

114

115 116

117

118

119

120

121

123

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138 139

140

141

142

Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

115

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

Optimal Human Embryo Culture

Optimal Human Embryo Culture

143

144

145

146

147

148 149

150

151

152

153

154

155

156

157

158

159

160

161

162

Swain

and glutathione status. In Vitro Cell Dev Biol Anim 1998;34(8): 655–659 Hong K, Forman E, Lee H, et al. Optimizing the temperature for embryo culture in IVF: a randomized controlled trial (RCT) comparing standard culture temperature of 37C to the reduced more physiologic temperature of 36C. Fertil Steril 2012;98(3):s167 Cortezzi SS, Garcia JS, Ferreira CR, et al. Secretome of the preimplantation human embryo by bottom-up label-free proteomics. Anal Bioanal Chem 2011;401(4):1331–1339 Beardsley AJ, Li Y, O’Neill C. Characterization of a diverse secretome generated by the mouse preimplantation embryo in vitro. Reprod Biol Endocrinol 2010;8:71 Brison DR, Hollywood K, Arnesen R, Goodacre R. Predicting human embryo viability: the road to non-invasive analysis of the secretome using metabolic footprinting. Reprod Biomed Online 2007;15(3):296–302 Katz-Jaffe MG, Schoolcraft WB, Gardner DK. Analysis of protein expression (secretome) by human and mouse preimplantation embryos. Fertil Steril 2006;86(3):678–685 Bormann C, Swain J, Ni Q, et al. Preimplantation embryo secretome identification. Fertil Steril 2006;86(Suppl 2):s116 Trimarchi JR, Liu L, Smith PJ, Keefe DL. Noninvasive measurement of potassium efflux as an early indicator of cell death in mouse embryos. Biol Reprod 2000;63(3):851–857 Trimarchi JR, Liu L, Porterfield DM, Smith PJ, Keefe DL. A noninvasive method for measuring preimplantation embryo physiology. Zygote 2000;8(1):15–24 Trimarchi JR, Liu L, Porterfield DM, Smith PJ, Keefe DL. Oxidative phosphorylation-dependent and -independent oxygen consumption by individual preimplantation mouse embryos. Biol Reprod 2000;62(6):1866–1874 Canseco RS, Sparks AE, Pearson RE, Gwazdauskas FC. Embryo density and medium volume effects on early murine embryo development. J Assist Reprod Genet 1992;9(5):454–457 Lane M, Gardner DK. Effect of incubation volume and embryo density on the development and viability of mouse embryos in vitro. Hum Reprod 1992;7(4):558–562 Kato Y, Tsunoda Y. Effects of the culture density of mouse zygotes on the development in vitro and in vivo. Theriogenology 1994; 41(6):1315–1322 Salahuddin S, Ookutsu S, Goto K, Nakanishi Y, Nagata Y. Effects of embryo density and co-culture of unfertilized oocytes on embryonic development of in-vitro fertilized mouse embryos. Hum Reprod 1995;10(9):2382–2385 Teruel M, Smith R. Effect of embryo density and growth factors on in vitro preimplantation development of mouse embryos. Acta Physiol Pharmacol Ther Latinoam 1997;47(2):87–96 Gardner DK, Lane M, Spitzer A, Batt PA. Enhanced rates of cleavage and development for sheep zygotes cultured to the blastocyst stage in vitro in the absence of serum and somatic cells: amino acids, vitamins, and culturing embryos in groups stimulate development. Biol Reprod 1994;50(2):390–400 Donnay I, Van Langendonckt A, Auquier P, et al. Effects of coculture and embryo number on the in vitro development of bovine embryos. Theriogenology 1997;47(8):1549–1561 O’Doherty EM, Wade MG, Hill JL, Boland MP. Effects of culturing bovine oocytes either singly or in groups on development to blastocysts. Theriogenology 1997;48(1):161–169 Fujita T, Umeki H, Shimura H, Kugumiya K, Shiga K. Effect of group culture and embryo-culture conditioned medium on development of bovine embryos. J Reprod Dev 2006;52(1):137–142 Keefer CL, Stice SL, Paprocki AM, Golueke P. In vitro culture of bovine IVM-IVF embryos: Cooperative interaction among embryos and the role of growth factors. Theriogenology 1994;41(6): 1323–1331 Khurana NK, Niemann H. Effects of oocyte quality, oxygen tension, embryo density, cumulus cells and energy substrates

Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

163

164

165

166

167

168

169

170

171

172

173

174

175

176

177

178

179

180

181

182

on cleavage and morula/blastocyst formation of bovine embryos. Theriogenology 2000;54(5):741–756 Nagao Y, Iijima R, Saeki K. Interaction between embryos and culture conditions during in vitro development of bovine early embryos. Zygote 2008;16(2):127–133 Larson MA, Kubisch HM. The effects of group size on development and interferon-tau secretion by in-vitro fertilized and cultured bovine blastocysts. Hum Reprod 1999;14(8):2075–2079 Reed M, Woodward B, Swain J. Single versus group culture of mammalian embryos: the verdict of the literature. J Reprod Stem Cel Biol 2011;2(2):77–87 Moessner J, Dodson WC. The quality of human embryo growth is improved when embryos are cultured in groups rather than separately. Fertil Steril 1995;64(5):1034–1035 Almagor M, Bejar C, Kafka I, Yaffe H. Pregnancy rates after communal growth of preimplantation human embryos in vitro. Fertil Steril 1996;66(3):394–397 Rebollar-Lazaro I, Matson P. The culture of human cleavage stage embryos alone or in groups: effect upon blastocyst utilization rates and implantation. Reprod Biol 2010;10(3):227–234 Spyropoulou I, Karamalegos C, Bolton VN. A prospective randomized study comparing the outcome of in-vitro fertilization and embryo transfer following culture of human embryos individually or in groups before embryo transfer on day 2. Hum Reprod 1999;14(1):76–79 Ebner T, Shebl O, Moser M, Mayer RB, Arzt W, Tews G. Group culture of human zygotes is superior to individual culture in terms of blastulation, implantation and life birth. Reprod Biomed Online 2010;21(6):762–768 Smith GD, Takayama S, Swain JE. Rethinking in vitro embryo culture: new developments in culture platforms and potential to improve assisted reproductive technologies. Biol Reprod 2012; 86(3):62 Swain JE, Lai D, Takayama S, Smith GD. Thinking big by thinking small: application of microfluidic technology to improve ART. Lab Chip 2013;13(7):1213–1224 Swain JE, Smith GD. Advances in embryo culture platforms: novel approaches to improve preimplantation embryo development through modifications of the microenvironment. Hum Reprod Update 2011;17(4):541–557 Vajta G, Korösi T, Du Y, et al. The well-of-the-well system: an efficient approach to improve embryo development. Reprod Biomed Online 2008;17(1):73–81 Morbeck DE. Importance of supply integrity for in vitro fertilization and embryo culture. Semin Reprod Med 2012;30(3): 182–190 Lee BE, Boone WR, Brackelsberg PO, Carmichael RA. Development of screening systems for evaluation of materials used in mammalian embryo transfer. Theriogenology 1988;30(3):605–612 Scott LF, Sundaram SG, Smith S. The relevance and use of mouse embryo bioassays for quality control in an assisted reproductive technology program. Fertil Steril 1993;60(3):559–568 Esterhuizen AD, Bosman E, Botes AD, et al. A comparative study on the diagnostic sensitivity of rodent sperm and embryos in the detection of endotoxin in Earle’s balanced salt solution. J Assist Reprod Genet 1994;11(1):38–42 Rinehart JS, Bavister BD, Gerrity M. Quality control in the in vitro fertilization laboratory: comparison of bioassay systems for water quality. J In Vitro Fert Embryo Transf 1988;5(6):335–342 van den Bergh M, Baszó I, Biramane J, Bertrand E, Devreker F, Englert Y. Quality control in IVF with mouse bioassays: a four years’ experience. J Assist Reprod Genet 1996;13(9):733–738 Gardner DK, Reed L, Linck D, Sheehan C, Lane M. Quality control in human in vitro fertilization. Semin Reprod Med 2005;23(4): 319–324 Punt-van der Zalm JP, Hendriks JC, Westphal JR, Kremer JA, Teerenstra S, Wetzels AM. Toxicity testing of human assisted

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

116

Optimal Human Embryo Culture

117

185 Davidson A, Vermesh M, Lobo RA, Paulson RJ. Mouse embryo

culture as quality control for human in vitro fertilization: the onecell versus the two-cell model. Fertil Steril 1988;49(3):516–521 186 Khan Z, Wolff HS, Fredrickson JR, Walker DL, Daftary GS, Morbeck DE. Mouse strain and quality control testing: improved sensitivity of the mouse embryo assay with embryos from outbred mice. Fertil Steril 2013;99(3):847–854.e2 187 Wolff HS, Fredrickson JR, Walker DL, Morbeck DE. Advances in quality control: mouse embryo morphokinetics are sensitive markers of in vitro stress. Hum Reprod 2013;28(7):1776–1782

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

reproduction devices using the mouse embryo assay. Reprod Biomed Online 2009;18(4):529–535 183 Morimoto Y, Hayashi E, Ohno T, Kawata A, Horikoshi Y, Kanzaki H. Quality control of human IVF/ICSI program using endotoxin measurement and sperm survival test. Hum Cell 1997;10(4): 271–276 184 Fleetham JA, Pattinson HA, Mortimer D. The mouse embryo culture system: improving the sensitivity for use as a quality control assay for human in vitro fertilization. Fertil Steril 1993; 59(1):192–196

Swain

Seminars in Reproductive Medicine

Vol. 33

No. 2/2015

Copyright of Seminars in Reproductive Medicine is the property of Thieme Medical Publishing Inc. and its content may not be copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written permission. However, users may print, download, or email articles for individual use.

Optimal human embryo culture.

A large contributor to success during in vitro fertilization (IVF) lies in the processes occurring within the IVF laboratory. These processes make up ...
331KB Sizes 5 Downloads 13 Views